Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Integrated cancer tissue engineering models for precision medicine

  • Michael E. Bregenzer ,

    Contributed equally to this work with: Michael E. Bregenzer, Eric N. Horst, Pooja Mehta, Caymen M. Novak, Shreya Raghavan, Catherine S. Snyder

    Affiliation Department of Biomedical Engineering, University of Michigan, Ann Arbor, Michigan, United States of America

  • Eric N. Horst ,

    Contributed equally to this work with: Michael E. Bregenzer, Eric N. Horst, Pooja Mehta, Caymen M. Novak, Shreya Raghavan, Catherine S. Snyder

    Affiliation Department of Biomedical Engineering, University of Michigan, Ann Arbor, Michigan, United States of America

  • Pooja Mehta ,

    Contributed equally to this work with: Michael E. Bregenzer, Eric N. Horst, Pooja Mehta, Caymen M. Novak, Shreya Raghavan, Catherine S. Snyder

    Affiliation Department of Materials Science and Engineering, University of Michigan, Ann Arbor, Michigan, United States of America

  • Caymen M. Novak ,

    Contributed equally to this work with: Michael E. Bregenzer, Eric N. Horst, Pooja Mehta, Caymen M. Novak, Shreya Raghavan, Catherine S. Snyder

    Affiliation Department of Biomedical Engineering, University of Michigan, Ann Arbor, Michigan, United States of America

  • Shreya Raghavan ,

    Contributed equally to this work with: Michael E. Bregenzer, Eric N. Horst, Pooja Mehta, Caymen M. Novak, Shreya Raghavan, Catherine S. Snyder

    Affiliation Department of Materials Science and Engineering, University of Michigan, Ann Arbor, Michigan, United States of America

  • Catherine S. Snyder ,

    Contributed equally to this work with: Michael E. Bregenzer, Eric N. Horst, Pooja Mehta, Caymen M. Novak, Shreya Raghavan, Catherine S. Snyder

    Affiliation Department of Materials Science and Engineering, University of Michigan, Ann Arbor, Michigan, United States of America

  • Geeta Mehta

    mehtagee@umich.edu

    Affiliations Department of Biomedical Engineering, University of Michigan, Ann Arbor, Michigan, United States of America, Department of Materials Science and Engineering, University of Michigan, Ann Arbor, Michigan, United States of America, Rogel Cancer Center, School of Medicine, University of Michigan, Ann Arbor, Michigan, United States of America, Macromolecular Science and Engineering, University of Michigan, Ann Arbor, Michigan, United States of America

Abstract

Tumors are not merely cancerous cells that undergo mindless proliferation. Rather, they are highly organized and interconnected organ systems. Tumor cells reside in complex microenvironments in which they are subjected to a variety of physical and chemical stimuli that influence cell behavior and ultimately the progression and maintenance of the tumor. As cancer bioengineers, it is our responsibility to create physiologic models that enable accurate understanding of the multi-dimensional structure, organization, and complex relationships in diverse tumor microenvironments. Such models can greatly expedite clinical discovery and translation by closely replicating the physiological conditions while maintaining high tunability and control of extrinsic factors. In this review, we discuss the current models that target key aspects of the tumor microenvironment and their role in cancer progression. In order to address sources of experimental variation and model limitations, we also make recommendations for methods to improve overall physiologic reproducibility, experimental repeatability, and rigor within the field. Improvements can be made through an enhanced emphasis on mathematical modeling, standardized in vitro model characterization, transparent reporting of methodologies, and designing experiments with physiological metrics. Taken together these considerations will enhance the relevance of in vitro tumor models, biological understanding, and accelerate treatment exploration ultimately leading to improved clinical outcomes. Moreover, the development of robust, user-friendly models that integrate important stimuli will allow for the in-depth study of tumors as they undergo progression from non-transformed primary cells to metastatic disease and facilitate translation to a wide variety of biological and clinical studies.

Introduction

Tumors have long been viewed as the accumulation of a mass of aberrant cancer cells. However, research has repeatedly shown the dependence of cancer progression on a variety of environmental factors, including non-cancerous cells, mechanical stimuli, and the surrounding extracellular matrix (ECM), aptly naming it as a ‘cancer-organ’. Although many in vitro and computational models currently exist, the complex and interdependent microenvironmental regulation of the ‘cancer-organ’ system at the dynamic tissue and molecular scale have not been fully addressed.

Tumorigenesis and cancer formation is a complex multistep process involving genetic, epigenetic, and metabolic alterations, and interactions with the microenvironment that transform normal cells into malignant ones. As part of this process, oncogenes get activated, and tumor suppressor genes get repressed, affecting cell proliferation, apoptosis, pro-tumoral inflammation, avoiding immune surveillance and destruction, promoting genomic instability, angiogenesis, and metastasis[1,2].

As the tumors progress, new aberrant blood vessels continue to sprout due to activation of angiogenic switches in order to sustain proliferating malignant cells. The excessively proliferating autonomous neoplastic cells invade the local tissue, following which they intravasate into nearby blood and lymphatic vessels. Through these conduits, the disseminated cancer cells transit to distant organs, ultimately homing into specific niches after extravasating the blood/lymph vessel lumima. At the secondary sites, they form micrometastasis, which include small nodules of cancer cells, followed by growth of these lesions into macroscopic tumors, leading to metastatic colonization[1,2].

Due to diverse interactions involved, cancers are highly heterogeneous organ-like masses. Their complex microenvironments not only contain the tumor cells, but also various infiltrating endothelial, hematopoietic, stromal, immune and other cell types, ECM components, biophysical characteristics and mechanical stimuli [35]. Interactions within microenvironment also help create metabolic changes, such as a hypoxic environment and nutrient fluctuations, which further contribute to heterogeneity of cancer cells.

With this multifaceted network of communication between the native tissue and the tumor taken into consideration, cancer is more aptly understood as a complex organ, dependent on and working within the various colonized organs. This view of cancer provides a realistic perspective which allows us to increase our understanding of the disease, and thus identify crucial aspects for facilitating drug screening and development of efficacious, individualized cancer therapies.

Investigative approaches and interpretation of the ‘cancer-organ’ system heavily influences research conclusions. For example, the growth of cells on 2-dimensional (2D) surfaces versus 3-dimensional (3D) constructs alters a cancer cell’s response to chemotherapeutics, thus influencing drug development and perceived effectiveness[6]. Similarly, mechanical stimuli innate to the microenvironment and exacerbated by the growth and development of the tumor can alter the stemness of the cancer cells[7] along with metastatic tendencies[810]. Meanwhile, cellular interactions between the non-malignant cell populations, immune components[11,12], and cancer cells influence the advancement of the disease, as well as, the response to common treatments[13]. Additionally, acellular aspects of the microenvironment, including soluble signaling and ECM composition and architecture, play a large role in phenotypic behavior[14,15] and thus the conclusions of experimental outcomes. Each of these factors uniquely impacts cellular components within the tumor microenvironment (TME), contributing to the complexity of the ‘cancer-organ’ system (Fig 1). However, our in depth understanding of these factors and their complex interplay is limited by current model systems, which fail to corroborate findings and elicit sufficient reproducibility within the field.

thumbnail
Fig 1. Components of the ‘Cancer-Organ’ model.

To develop an accurate multi-dimensional understanding of the structure, organization, and complex relationships in cancers, we need to consider the following factors. Heterogeneous cancer cells reside in a complex tumor microenvironment, which consists of mechanical stimuli, non-malignant cell-cancer cell interactions, soluble signals, and extracellular matrix (ECM). The dimensionality of cell culture influences cancer cell motility and cellular interaction with the surrounding cells and ECM. Mechanical stimuli including shear, compressive, tensile, and viscoelastic forces, dynamically influence cancer cells as the tumor grows. Similarly, cellular interactions through direct contact with surrounding non-malignant cells and soluble signals alter communication and downstream signaling. Interactions between immune cells and cancerous cells are highly complex and can lead to immune evasion and support of tumor progression. All of these characteristics play an integral role in tumor progression and are critical to forming a complete picture of the ‘cancer-organ’ system.

https://doi.org/10.1371/journal.pone.0216564.g001

Therefore, the question remains, how do we as researchers reframe our understanding of cancer to encompass the many key players within the system. To address this dilemma, we have compiled a review of impactful cancer bioengineering models that investigate the important factors within the ‘cancer-organ’ system, including: dimensionality of cell culture (Section I), mechanical stimuli (Section II), multicellular interactions (Section III), immune interactions (Section IV), soluble signaling (Section V), complex cancer bioengineering models (Section VI), and mathematical models (Section VII), (Fig 1). These model systems portray the application of tissue engineering principles in understanding cancer biology and translating discoveries to delivery systems and precision medicine. We shed light on current state-of-the-art engineering methodologies that can help construct integrative cancer models in the first part of the review (Fig 2). In the latter half of the review, we provide suggestions of how to improve the quality and reproducibility of in vitro models and their findings (Section VIII), enabling the accelerated progress of cancer research as a whole. The ultimate goal of these integrated multi-scale models is not only to improve the understanding of cancer biology, but also to catalyze effective and personalized drug screening and therapeutic strategies that take the entire integrated complex TME into consideration.

thumbnail
Fig 2. Various engineering tools can help construct the complex picture of the ‘cancer-organ’ system.

Summarized here are the state-of-the-art cancer bioengineering models that we discuss in this review. Each model has inherent benefits and drawbacks that are discussed in more detail within the following sections. We have listed the components of the ‘cancer-organ’ system which can be probed with the specific model in the figure.

https://doi.org/10.1371/journal.pone.0216564.g002

Dimensionality of the ‘Cancer-Organ’ models

Integrated ‘cancer-organ’ in vitro models are limited by the continued use of 2D cell culture, common due to ease of use. However, cancers share no similarity to cells grown in 2D, and as previously described above, grow as ‘organs’[35]. To improve the biologic relevance of ‘cancer-organ’ models, many studies have shown that 3D cultures are more reflective of the in vivo TME resulting in more physiologic cell behavior[6,16]. Specifically, compared to cells cultured in flat, rigid 2D culture dishes, cells in 3D cultures have different spatial arrangement of their surface receptors because they are surrounded by other cells and have 3D spatial constraints[17]. This 3D arrangement ultimately alters cell polarity, signal transduction, gene expression, cell morphology, growth rates, and other phenotypes[1821]. 3D culture is also inherently more representative of the in vivo TME, as it allows for 3D nutrient and oxygen gradients, as compared to 2D culture, which results in a homogeneous distribution of nutrients and oxygen[22,23]. 2D cultures may also skew experimental results through unintentional selection of proliferating cells, as necrotic cells will not adhere to the tissue culture dish and will be removed during standard cell culture maintenance[23]. As a result of these differences, research has shown that drug response found in 3D models is more reflective of an in vivo response as compared to 2D cultures.

Consequently, growing number of cancer biologists and engineers alike are exploring 3D culture methods, as they have been shown to be more representative of the in vivo TME than traditional 2D methods[2428]. For example, when comparing lung cancer cell growth and behavior on a polyester-based composite 3D scaffold versus 2D monolayer culture, it was found that the cancer cells had morphology more representative of in vivo tumors including necrotic centers, as well as upregulation of CD44 and carbonic anhydrase IX[29]. MCF10A breast cancer cells demonstrated changes in IL-6, H-Ras and E-Cadherin expression when grown in 3D within a Matrigel hydrogel, similar to those found in vivo [30]. Additionally, breast cancer cells were shown to have enhanced HER2 activation in 2D culture as well as a switch in signaling from phosphoinositide 3-kinase in 2D culture to mitogen-activated protein kinase in a 3D cell spheroid[31]. Drawing from these findings, it is clear the dimensionality of cancer cell culture greatly influences cell phenotypes, protein expression, drug response, precision medicine, and thus experimental results. The methods and materials chosen to formulate the 3D environment will interact uniquely with cell cultures, and these inherent benefits and drawbacks of each are discussed below.

The currently available 3D cell culture models include non-adherent suspension culture, hydrogels, 3D bioprinting, and scaffolds. Within each of these subtypes, material and fabrication variations can alter important culture parameters, such as cell adhesion, ECM structure, and ECM stiffness. In non-adherent 3D cultures, all possible cell attachment structures are removed as with rotary vessel/spinner flasks[32], hanging drop arrays[32,33], superhydrophobic surfaces[34], aqueous two-phase systems[35], and liquid overlays[36]. These culture methods usually result in the formation of spheroids, 3D microtissue aggregates of cells in the culture. These non-adherent models by their nature emphasize cell-cell interactions and are discussed in detail in Section III.I.

Another method of creating a 3D culture is to introduce an ECM-like element, such as a hydrogel or scaffold. These ECM mimics can fully encapsulate the cells within the matrix material, providing cells with structural, adhesive, mechanical and physical cues. Adherent hydrogels provide 3D support by allowing the cells to move and interact in all three (x, y and z) directions and interface with their environment on all surfaces of the cell as opposed to only laterally (x and y) in a 2D culture. Hydrogels also tend to emphasize cell-matrix interactions, which also include mechanical stimuli. Scaffolds are typically more rigid than hydrogels, and have micro- or macro-porous structures. They allow cells to migrate in 3D, interact with other cells in the construct, and provide a spacious microstructure that allows for cell-based ECM construction. Cell-matrix and cell-cell interactions are present due to the cell clusters that grow within the scaffolds. The chemical composition of the material chosen is crucial as it influences cellular response in terms of proliferation, migration, matrix rearrangement, and cell cluster formation[37,38]. For example, a biologically recognizable material such as collagen stimulates cell adhesion while cells do not interact with an inert material such as agarose[39,40], which acts as a blank background environment. Scaffolds and hydrogels can also be combined, creating a unique 3D environment that can both encapsulate cells while providing enhanced rigidity of the overall construct.

Of note, cells cultured in a 3D scaffold show more similarities to cells in human tumor tissues compared to 2D cultures. For example, glioblastoma U87 cells form round or ovoid shapes, and develop complex structures with cilia or microvilli when grown on a 3D collagen scaffold but exhibit an epithelioid morphology in 2D culture. These cells also exhibited therapeutic responses similar to responses seen in patients with glioblastoma[41]. Other properties of 3D cultures, such as stiffness and porosity, can also be adjusted through choosing or tuning an appropriate material or fabrication process to influence cell viability, migration capability, and flow of nutrients and waste in the culture[38]. A major limitation of most in vitro 3D culture systems is the lack of vasculature, which may affect our understanding of drug efficacy due to its role in drug, oxygen, and nutrient delivery. Though the 3D system does recreate cell-cell and cell-ECM interactions, it remains difficult to introduce physiologically relevant vascularization into these constructs, limiting the nutrient and waste exchange to the cell cultures, though some systems have been successful[42,43]. The challenge of vascularity models is further discussed in Sections V and VI.

The 3D culture method chosen for a given experiment should be chosen carefully based on the biological question being asked and on the inherent benefits and drawbacks, summarized in Table 1. Conscious design of 3D cultures with regard to these parameters is imperative due to their effects on cell behavior and overall effect of the ‘cancer-organ’ system, highlighting the need to develop model standards that most accurately replicate in vivo conditions. 3D models still currently lack a standardized reporting and characterization system. Without these standards, our understanding of the complexity of the ‘cancer-organ’ system falls short and creates the conditions for discrepancies between findings and lack of reproducibility. Suggestions for how to standardize reporting of 3D culture properties are discussed in section VII.II.

thumbnail
Table 1. Summary of 3-dimensional cancer bioengineering methods and their respective benefits and limitations.

https://doi.org/10.1371/journal.pone.0216564.t001

Mechanical Stimuli in the ‘Cancer-Organ’ models

The tumor microenvironment in the ‘cancer-organ’ induces mechanical tension, compression, and shear stress on the growing tumor, while exposing the cells to increased ECM stiffness and variable viscoelasticity. Therefore, these stimuli are explored in many mechanical cancer models[10,6769]. For samples of micro an macro bioreactors utilized for examining mechanical stimuli in cancer, readers are referred to references [7,7073]. As a wide variety of bioreactors exists to explore the dynamic mechano-environment a comprehensive list of devices is beyond the scope of this review. Over-proliferative tumor cells and an increase in interstitial fluid pressure, caused by tumor-initiated angiogenesis, orchestrate circumferential mechanical stretching along the tumor’s leading edge. To study this phenomenon, microreactors with flexible membranes are often used to stimulate stretching within the TME. Stretching using the microreactor models has been shown to promote cancer cell growth and induce proliferation[70,74,75], as well as upregulation of the YAP/TAZ pathways[74]. Commercially available bioreactors provide varying levels of uniaxial or equiaxial tensile force, and can be applied to ‘cancer-organ’ models[76].

The surrounding tissue provides resistance to the expanding tumor. As a consequence, the tumor is exposed to high levels of solid stress and the cancer cells experience ever increasing compressive force[7779]. Scaffolds to study cancer compressive mechanotransduction have been fabricated using poly(lactide-co-glycolide) or hyaluronic acid, seeded with cells, and exposed to cyclic loading via compression bioreactors[80,81]. These bioreactors are typically designed and built in-house to allow for fine-tuned compressive loading cycles, although commercial options do exist. As a variation to this approach, a hydrogel can be embedded with cells and exposed to static compression by using a weight or piston to achieve the desired force[9,73,82,83].

The role of TME stiffness on cancer phenotype has been studied with a variety of models. For example, surface functionalized PDMS microposts have been engineered for specific stiffness and used to evaluate individual cell mechanics and protein expressions[84]. In another approach, increasing the polymer concentration of hydrogels during fabrication also modulates the hydrogel’s stiffness. Employing this technique allows the effects of stiffness to be tested on cancer cells without changing the substrate to which cells adhere[71,8587]. Optical tweezers have also been used to study the effects of stiffness of single cancer cells[88]. Dynamic ECM models that replicate the ECM remodeling during cancer progression to support tumor growth, are increasingly becoming popular, since they modulate physical properties over time[89].

Most human tissue and polymer or protein ECM analogues experience varying degrees of an elastic strain during loading cycles. These viscoelastic matrices show time dependent recovery when loads are removed. In recent years, the viscoelastic nature of the TME has been shown to impact tumor matrix remodeling in collagen, fibrin, alginate, reconstituted basement membrane, and agarose hydrogel models[90]. Viscoelasticity has also been shown to impact cancer cell invasion in interpenetrating network hydrogels with low molecular weight RGD-alginate and reconstituted basement membrane, as well as collagen type I hydrogels[91,92]. Given the importance of ECM viscoelasticity in cancer progression and metastasis, additional studies are required to model and probe viscoelastic changes in the TME.

In addition, the TME is under a constant barrage of fluid-induced shear stress. Leaky vasculature within the tumor niche as well as venous blood flow has been shown to exert shear stresses ranging from 0.5 to 4.0 dyn/cm2. Circulating tumor cells and metastatic cells undergoing intravasation and extravasation may also experience a range of arterial shear stress from 4 to 30 dyn/cm2[93]. Shear stress is often tested using a microfluidic device design where growth medium is pumped through the closed system using a syringe or circulating pump. A narrowing of the flow channel within the device allows for pronounced wall shear stress and controlled laminar flow. We refer the reader to the review by Huo et al. for other forms of microfluidic devices implemented in the field of cancer mechanobiology[94].

Shear stress stimulation has been shown to increase proliferation[72], upregulate the pro-survival ERK pathways[95], enhance motility via YAP/TAZ[96], and increase chemoresistance[7]. Parallel plate[97] and rotary bioreactors[98] that apply shear stress to cells have also been used to study adhesion mechanics of tumor cells. Perfusion bioreactors typically provide cancer cells with relatively uniform shear stress across the entirety of the polymer scaffold. The TME may also be fine-tuned by manipulating the composition of the polymer or hydrogel. Taken together, there is ample evidence of the significant role that the mechanical forces in the TME play in tumor progression and of the variability present between models used to study mechanical stimuli. Importantly, many bioengineered models fail to consider mechanical stimuli all together, potentially resulting in unrealistic results. It is clear that mechanical stimuli influence key hallmarks of cancer and that the tumor requires this stimulus to elicit specific functionality; however, how cellular processes sense each of these stresses and translate them to oncogene expression is still poorly understood.

Multicellular interactions in the ‘Cancer-Organ’ models

Cells and ECM come together to form tissues, which then collectively form structurally stable and functional organs. In contrast, tumors are a non-random mix of cells and ECM but are functionally and structurally unstable and abnormal. The cues from the TME non-cancerous cell types (including immune cells, endothelial cells, fibroblasts, and mesenchymal stem cells, among many others) have been shown to be instrumental in tumor initiation, progression, and metastasis[99102]. In this section and in Fig 3, we describe some of the important modes of cell-cell communication components and their roles in the TME and ‘cancer-organ’ models.

thumbnail
Fig 3. Various cell-cell interactions within the cancer-organ system.

Interactions of cancer and malignant cells with their surroundings help dictate their survival and phenotypes. Within the homeostatic non-transformed microenvironment, various cell-cell junctions are formed ensuring the proper polarization, orientation, and proliferation of the non-malignant cells. Cell-ECM interactions provide structure and mechanical stimuli to the cellular surroundings through points of adhesion. These native interactions are disrupted by the infiltrating cancer cells which interrupt cell-cell communications and displace healthy tissue. The cancer cells undergo the epithelial-mesenchymal transition in order to metastasize and do not experience the same proliferative inhibition provided by non-malignant cell-cell communication. Well-established communication between cancerous cells increases survival by avoiding anoikis and promoting chemoresistance. Finally, the surrounding ECM, which is stiffened by the presence of the expanding cancer mass, aids in additional cancer cell migration, and an altered mechanical environment will feed forward the progression of the disease.

https://doi.org/10.1371/journal.pone.0216564.g003

Cells within the TME communicate via soluble signals, as well as, through physical connections, such as communication junctions (connexins, ion channels, chemical synapses, and pannexins), occluding junctions (tight junctions), and anchoring junctions (adherens, desmosomes, focal adhesions, and hemidesmosomes)[72]. These physical cell-cell communication components, which comprise specialized intercellular junctional proteins, are critical for maintaining cell polarity, barrier function, morphogenesis, differentiation, homeostasis, cell growth, and cell-cell interactions. All of these junctional complexes are used by cancer cells to transmit signals to the neighboring cells, as well as, to respond cohesively to various conditions. Furthermore, integrins facilitate cell-ECM interactions via focal adhesions and hemidesmosomes[103107], allowing for cells to communicate with their physical surroundings.

These junctions are often disrupted in cancers via genetic mutation or epigenetic changes ultimately affecting tumor progression[108]. For example, tight junctions create a barrier in endothelial cells, allowing molecules and inflammatory cells to pass, whereas within epithelial cells, tight junctions work in an adhesive manner, keeping cells correctly polarized and preventing cells from distortion[109,110]. During intravasation, cancer cells distort the tight junctions of the vascular endothelium to penetrate it, which is considered one of the most critical steps in the dissemination of cancer cells. Thus, the tight junctions are the initial barriers the cancer cells must overcome in order to metastasize[111,112]. Similarly, an important step in metastasis is the epithelial-to-mesenchymal transition (EMT), which involves a cadherin switch from E-cadherin to N-cadherin promoting the migratory capacity of the tumor cells[108].

Additionally, previous studies have indicated that desmosomes undergo transformation during cancer progression[113]. Studies have shown that desmosomal proteins have both tumor-promoting[114] and tumor-suppressive[113] functions in different types of cancers. In particular, desmocollin1 (DSC1) and desmocollin3 (DSC3) may act as prognostic markers for lung cancer[115], colorectal cancer[113] and in esophageal, head and neck cancers[116]. As a major component of sensing the mechanical environment and the surrounding cells, proteins involved in cell adhesion are crucial aspects of tumor development. Cancer bioengineering models need to mimic these physiologic cell-cell and cell-matrix adhesions, interactions, and modifications in cancer to remain true to the complexity of cancer. A summary of these interactions is depicted in Fig 3.

3.1 ‘Cancer-Organ’ models emphasizing direct cell-cell interactions

Among the currently available 3D physiological integrated models of cancers, the following prioritize cell-cell interactions: 1) multicellular tumor spheroids; 2) tumorospheres; 3) tissue-derived tumor spheres; 4) organotypic multicellular spheroids[117,118] and; 5) organoids[119,120]. We described these 3D models in Section I. Others include microfluidics, aqueous two-phase system, and microfabricated microwell array[118,121]. The following section details the features of and differences between these 3D cancer models. Table 2 highlights some of these examples.

thumbnail
Table 2. Examples of 3-dimensional cancer bioengineering models that emphasize cell-cell interactions.

https://doi.org/10.1371/journal.pone.0216564.t002

Multicellular tumor spheroids are generated from suspensions of single cells of immortalized cell lines in the presence of serum in non-adherent conditions[64,65,121,122]. Originally formed with just cancer cells, multicellular tumor spheroids have more recently been cultured with combinations of cancer cells with immune cells, fibroblasts, and endothelial cells to study heterogeneous interactions in the tumor tissue. These spheroids can vary in diameter from 100’s of microns to 3mm, with a degree of compaction depending on the cell line of origin and culture method[32,123125]. At diameters larger than 400μm, spheroids can have an inner core of hypoxic quiescent cells and an outer layer of proliferating cells replicating the hypoxic pockets that can form within tumors[117]. They can also replicate the differentiation of the parent tumor and have been demonstrated to enrich for cancer stem-like cells[126]. Studies have shown multicellular tumor spheroids are more physiologically representative and display similarities to patient tumors in not only their proliferative index but also to cell morphology, cell–cell junctions, and ERK1/2, MAPK, and PI3K, AKT pathway activation[127].

Patient tumor-derived tumorospheres are obtained from fine slicing and partial dissociation of cancer tissue. These tumorospheres represent the histological features, gene expression profiles, mutations, and tumorigenicity of the parent tissue. Because tissue-derived tumorospheres are formed due to dissociation, they are exclusively composed of cancer cells[123125,127]. Researchers have shown that E-cadherin is involved in cell-cell interactions in tissue-derived tumorospheres and E-cadherin/β-catenin complexes were shown to be tethered to the cytoskeleton. This organization has been demonstrated to strengthen cell–cell adhesion in other systems, suggesting that tissue-derived tumorospheres have strong inter- and intra-cellular interactions[124,127,128]. Tumorospheres can be cultured under serum free conditions from suspensions of single cells that are sorted from a population of cancer cells[128130]. Single cells then expand clonally to produce tumorospheres. Due to the capacity of stem cells to survive in serum free conditions and expand clonally, tumorospheres are specifically suited to cancer stem-like cells and their enrichment. In fact, mammosphere or neurosphere formation from a single cancer-initiating cell plated in suspension is a gold standard for tumor initiation research. While tumorospheres do not fully replicate the TME, they have been useful in understanding cancer stem-like cells[117].

Organotypic multicellular spheroids arise from slicing tumor tissue fragments into sub millimeter pieces followed by culturing in a non–adherent system with serum and other supplements[131133]. These spheroids are circular structures that can be frozen or cultured. Spheroids isolated from ovarian carcinoma ascites fluid are a special case in the organotypic multicellular spheroids family because unlike the other organotypic multicellular spheroids, they are not generated after tissue processing but are isolated directly from patient effusions. In addition to recapitulating the original heterogeneity of the tumor, these ovarian spheroids also maintain their stromal component[117,131]. For example, organotypic multicellular spheroids maintain the presence of macrophages and preserve vessels with striated fibers of collagen in association with fibroblasts that surround vascular elements[131]. In another example, organotypic multicellular spheroids from bladder cancer display cell cycle distribution, which is similar to that observed in original tumors[134]. For further analysis of the advantages and disadvantages of these four model types, we refer readers to a more thorough review by Weiswald et al.[117].

Organoids can be grown by embedding embryonic stem cells (ESCs), induced pluripotent stem cells (iPSCs), somatic stem cells, and cancer cells into a 3D matrix and letting the cells self-organize into ‘mini organs’ similar to the organ of origin in a serum free medium[120,135,136]. Apart from a 3D matrix, which acts as a substitute for ECM, most organoid cultures medium requires different growth factors to grow depending on the tissue of origin[137]. Organoids can be passaged serially every 1–2 weeks[136] and can be genetically manipulated with comparatively more ease than other cancer models[138,139]. Organoids have been formed with both healthy tissues, including kidney[140,141], lung[142], liver[143,144], brain[145], colon[146149] and from tumors derived tissues, such as those formed from breast cancer[137], prostate cancer[150], glioblastoma[151], pancreatic cancer[144,152154], liver cancer[144], and colon/colorectal cancer[146,155]. This biorepository of healthy and tumor-derived organoids is an extremely useful tool in studying drug screening, cancer development, and disease modeling and precision medicine[136].

Since cancer operates and presents as a complex organ, there are still gaps in knowledge that these models have not been able to fill despite their improved physiological relevance over traditional models. Although they have been made from multiple cancer types and have excellent reproducibility, they are not complete representations of the in vivo system, as most lack vascular networks and intact immune components. Additionally, these models do not allow for fine manipulation of ECM, as many of these models are non-adherent cultures, and the models that utilize Matrigel are subject to batch-to-batch variation. While addition of ECM substitutes, like Matrigel, provide for the lack of a basement membrane and may make the model more physiologically relevant, they may also incorporate undefined extrinsic factors[156] that may cause artificial experimental outcomes. Additionally, some of these spheroid models enrich for a rare population of cancer cells termed cancer stem-like cells but do not include surrounding non-malignant cells, thus preventing us from learning about the interactions of cancer stem-like cells with neighboring cells in their 3D niche. Furthermore, organoids derived from stem cells may be difficult to standardize due to the need for complex addition of soluble signals at precise temporal intervals to direct differentiation down the desired lineage[157]. Finally, the variability present in spheroid-based modeling techniques represents a barrier to consistency in cancer research findings, as the differences between each spheroid generation method can result in discrepancies in research findings[158]. Therefore, there is a need for more comprehensive models that incorporate the complex environment of cancer and better recapitulate cancer–cell communication and functioning of the ‘cancer-organ’.

3.2 Modulation of ratio and density of cells in ‘Cancer-Organ’ models

Regardless of the model utilized, the type and number of cells added to the ‘cancer-organ’ may vary depending on the stage and progression of cancer being studied. The ratio between differing cell types should also be considered to reflect the in vivo TME setting, considering the proliferation rate of each cell type and their properties when cultured in vitro[159]. In an example of this consideration, Eder et al. used the hanging drop method with a co-culture of prostate cancer cells and cancer-associated fibroblasts. They showed an increased number of cancer cells compared to fibroblasts within the spheroids, which reflects in vivo observations[160]. While non-cancerous cells initiate cell cycle arrest to stop proliferation, cancer cells proliferate indefinitely and display no contact inhibition. Expectedly, in a heterogenous 3D model of cancer, non-cancerous cells are subject to contact inhibition while cancer cells are not, thereby changing the ratio between cancer and non-cancerous cells and overall cell density[159]. Diffusion and exchange of soluble factors within cell culture is also directly influenced by heterogeneous cell density, emphasizing the need for spatial control of cell seeding. While most models do not allow for spatial control of cell seeding densities, 3D bioprinting methods for cancer cell patterning have recently been developed, including valve-based printing, laser based printing, and thermal and piezoelectric inkjet printing, which provide reproducible control over spatial distance between cell types (i.e., cancer and stromal cells)[161], theoretically enhancing reproducibility of model results. In summary, the seeding density, as well as, ratio between cell types needs to be carefully considered during the design of 3D integrated ‘cancer-organ’ models.

Immune interactions in the ‘Cancer-Organ’ models

The TME, although largely made up of neoplastic tumor cells, also contains stroma and several types of immune cells[162]. The immune contexture of solid tumors is very well described and reviewed for several different kinds of tumors[163,164]. Tumors feature infiltrating lymphocytes, mature and immature myeloid cells capable of differentiation, macrophages, dendritic cells, eosinophils, mast cells, natural killer cells, and myeloid derived suppressor cells[162,164]. The type, location, and density of immune cells within the tumor are considered valuable prognostic tools in the treatment of neoplastic malignancies[165168]. The general consensus for immune cells within the TME is that their dysregulated and somewhat functionally impaired phenotype results in an immunosuppressive TME, allowing for tumor progression[169173]. Alternatively, immune cells, such as activated macrophages in the TME, also play more direct roles in promoting angiogenesis and tumor cell migration and survival, leading to not just an immunosuppressive program but an overall thriving and conducive environment for tumor progression[169173]. In the cancer immunosurveillance paradigm, neoplasia is largely controlled in its initial stages by various immune cells, so much so that immune cells were clearly demonstrated to perform tumor-specific rejection in transplanted tumors in mice[174,175]. The crux of this hypothesis was the new antigenic properties of tumors elicited a potent immune response, leading to tumor regression. In fact, in our current understanding of immunologic escape by tumors despite immune surveillance, tumor variants with reduced immunogenicity inherently cull their high immunogenic counterparts, leading to tumor progression. Today, this concept is broadly described as ‘cancer immunoediting’ with a range of actions from anti- to pro-tumoral scenarios. The classic immune surveillance paradigm falls into the ‘Elimination’ phase of immunoediting. An equilibrium process is presented where low immunogenic tumor variants are selected. Lastly, an escape process is reached where the tumor actively suppresses and creates anergic versions of immune cells within the TME, thereby promoting tumor tolerance and tumor progression[176179]. A summary of the role of immune cells and their roles in the TME can be found in Fig 4.

thumbnail
Fig 4. The immune microenvironment of tumors contains cellular components from both the innate and adaptive immune systems, with functional immuno-modulation between all the different cell types.

Macrophages are typically the most abundant population of leukocytes within the TME, derived from both tissue-resident and circulating monocytic progenitors. The accumulation of tumor-associated macrophages is often correlated with the development of pathological phenotypes in cancer, which leads to the promotion of angiogenesis, metastasis, chemoresistance and functional suppression of adaptive immunity. The TME counterbalances activating natural killer (NK) cell signals with strong inhibitory signals to escape NK cell mediated immune surveillance and further reduce the phagocytic activity of NK cells. NK cells also exhibit functional anergic phenotypes with reduced phagocytosis and reduced amounts of cytoplasmic granules that contribute to tumor progression. Other granulocytes within the TME often recruited from circulating vasculature include neutrophils, basophils, eosinophils and mast cells. Tumors often experience reduced recruitment, but granulocytes are often re-programmed to a pro-tumor phenotype, promoting vascular normalization and stromal remodeling. Analysis of several solid tumors also indicate that they are infiltrated with T-cells and B-cells, recruited from circulating blood and lymphatic structures. The number of infiltrated T-cells offer significant prognostic value to cancers. However, the TME reprograms T-cells into an exhausted anergic state, leading to severe immune suppression, specifically of the Th and CTL (CD8+ cytotoxic T lymphocytes) phenotypes. Additionally, recruited naive T-cells are also converted to an insidious regulatory Treg phenotype, which contributes to suppressive immunomodulation. B-cells typically respond to tumor-derived antigens and elicit antibody responses through IgM secretion and direct stimulation of Th cells. Tumor-educated B-cells are immuno-suppressive, promote regulatory T-cells, and promote carcinogenesis. Myeloid derived suppressor cells are heterogeneous mixes of immature myeloid cells, found accumulated in lymphoid structures, blood, and the TME, and are heavily correlated with immune suppression. Myeloid derived suppressor cells are powerful inactivators of T-cells. Impaired myeloid differentiation also results in defective antigen presentation. Coupled with dysregulated T-cell priming by antigen presenters like dendritic cells, an overall immune suppressive landscape leads to tumor escape from immune surveillance.

https://doi.org/10.1371/journal.pone.0216564.g004

Until recently, most bioengineering models provided little insight into the role of the immune cells in the TME, with cancer research dominated by immunosuppressive in vivo models and mono-culture in vitro systems[180183]. Due to the important role that the immune cells can play in cancer progression, recent models have attempted to mimic the native immune component of the TME[184186]. Some of these models take advantage of patient’s peripheral blood mononuclear cells from which dendritic cells, lymphocytes, monocytes, or natural killer cells can be derived[187]. This is an important source of immune cells, as the derived cells will be representative of that specific patient’s immune system.

Outside of generating biomaterial strategies and testing on murine models, in vitro models also afford an easy platform for testing. In vitro systems are gaining increasing importance in the era of personalized medicine, where tumor antigens from the cancer genome will be identified on a personalized basis using deep genome sequencing techniques[188192]. For example, Herter et al. developed a heterotypic spheroid model, including tumor cells, fibroblasts, and immune cells. This system visualized immune cell infiltration and specific elimination of tumor cells upon immune cell activation with a novel immunocytokine IgG-IL2v and tumor or fibroblast targeted T-cell bispecific antibodies [193]. Another group created tumoroid and tumor slice cultures from patient-derived peripheral and tumor immune populations. This model demonstrated sensitivity and resistance to 5-fluorouracil and Lonsurf in the context of a heterogeneous tumor immune microenvironment[194]. The biggest advantage of these methods lies in the rapid establishment of the model system, which can be developed within weeks of surgical resection and provide personalized screening in the context of a functional immune system. The 3D architecture of such cancer/immune models helps to predict immunocytes interactions with tumor cells with higher fidelity. Indeed, culture of melanoma cells in 3D systems demonstrated impaired immuno-recognition by cytotoxic T lymphocytes when compared to 2D cultures[195]. Similarly, incorporating macrophages into 3D organotypic cultures of squamous cell carcinoma demonstrated a potent immuno-suppressive program, activating macrophages into a pro-tumoral alternatively activated phenotype with no external cytokine stimulation[196]. The inclusion of immune components in bioengineered in vitro models make them more representative of the in vivo tumor and will allow for improved cancer research and personalized treatments.

Recent progress in tumor immunology and identifying pharmaceutical targets focused on immunologic approaches have significantly moved the cancer immunotherapy field forward. While the idea of using immune tissue engineering to incorporate an immune component to cancer engineering has fundamental appeal, the immune system is a complex mesh of innate and adaptive immunity, with extensive functional immunomodulation between the two compartments. Facilitating this immunomodulation requires incorporation of several different types of immune cells, and sourcing these immune cells, whether from peripheral blood or tissue/tumor resident approaches, still requires better isolation procedures[4,197199]. Furthermore, a fundamental understanding of how the TME tips the balance of maintaining peripheral tolerance while suppressing adaptive immunity is imperative in engineering the immune tumor component.

Development of better engineered tumor-immune models will also accelerate the progression and impact of the emerging field of cellular immunotherapies. A big component of cellular immunotherapies began with the advent of genetically engineered T-cells with T-cell receptors engineered to direct their cytotoxic activity toward tumor cells. Chimeric Antigen Receptor T-cells (or CAR T-cells) are an emerging cellular immunotherapy paradigm to treat both solid and hematogenous tumors[200202]. CARs contain antigen recognition regions specifically directed against tumor-derived antigens or neo-antigens and intracellular domains that are combinations of co-stimulatory peptides. Engineered platforms can serve as wonderful tools to test and assess toxicity to mitigate adverse toxicity reactions reported in clinical trials for adoptive T-cell therapies[203,204]. For detailed reviews and perspectives on preclinical studies and early phase clinical trials in CAR T-cell therapy, readers are recommended to refer to Grigor et al. [205], Newick et al. [201], Mata et al. [206] and Yu et al. [207].

Other major challenges include the information not currently gained from a tumor’s genetic signature, including specifically the response of host immune cells within the TME. Several advances in enabling technologies are required for a high fidelity immune tissue engineering model that can replicate the immune complexity and response of the tumor when viewed as an organ. The following technologies in our opinion are of high value to translate tumor immune microenvironment engineering into patient-specific therapeutic strategies: rapid identification of the immune TME during diagnosis both locally within the tumor and peripherally incorporated from reliable clinical assays; integrated systems-based immuno-modeling approaches that could be used to develop predictive models based on individual immune cell behaviors[208210]; and microfluidics or small-cell based approaches to understand individual immune cell interactions (e.g. T-cell-macrophage, or T-cell-T-cell, etc.) to be integrated into a whole ‘cancer-organ’ perspective[211,212].

Soluble signaling in the ‘Cancer-Organ’ models

Just as soluble signals play an integral role in organ function and development[157], they are also critical in tumor development. In addition to direct cell-cell communication discussed earlier, cells within the TME also communicate through indirect means via soluble signaling and endocrine signaling[99101,103]. These soluble signals include growth factors[213,214], chemokines[214,215], cytokines[214,215], exosomes[214,215], and hormones[216]. Importantly, cells in the TME also respond to chemical gradients, which act as another form of soluble signal. Specifically, oxygen levels have proven to be key determiners of tumor progression[217]. The composition of these signals in the TME affects critical aspects of a tumor, such as angiogenesis[104,218220], proliferation[219], differentiation[221], drug response[220], and metastatic capacity[104,220,222]. In a tumor, soluble signal gradients can cause migration of immune cells[222], endothelial cells[219], and mesenchymal cells[223] to the tumor. These cells then act to either promote[224] or inhibit tumor growth[225]. Conversely, chemokine gradients may also contribute to the metastatic destination of tumor cells expressing the appropriate chemokine receptor[104]. Due to the complexity of soluble signaling in the TME, we will not discuss every facet in great detail, but rather we will highlight some key examples to further demonstrate the importance of soluble signaling in tumor progression and the models employed to study them.

A clear example of the profound effects that soluble signaling can have on tumor development is the effects of mesenchymal stem cell conditioned medium (MSC-CM) on tumor cell growth. A single dose of MSC-CM applied to SGC-7901 tumor cells resulted in tumor cell expression of VEGF and RhoA-GTPase and ERK1/2 activation leading to enhanced tumor growth that was maintained through serial transplantation experiments[224]. Furthermore, MSCs can differentiate into fibroblasts, which are typically responsible for synthesizing ECM components, maintaining tissue homeostasis, and regulating inflammation, proliferation, and differentiation in healthy tissue. In the presence of cancer cells, fibroblasts can be activated by cancer or immune secreted TGF-β, FGF-2, HGF, PDGF, and interleukins, as well as reactive oxygen species to become cancer-associated fibroblasts (CAFs)[223]. Activation of CAFs then results in an increased secretion of chemokines that promote tumor proliferation, invasion, and angiogenesis[223]. CAFs also interact with endothelial cells through secretion of FGF-2 and SDF-1. These factors promote angiogenesis and recruitment of endothelial cells respectively, which facilitate tumor growth and metastatic capability[219].

Additionally, long distance signaling with hormones (endocrine signaling) is an integral component of some cancers, so much so that therapeutics aiming to inhibit these signals have been developed[226,227]. For example, 17β-estrodiol (E2) is a hormone that affects the physiology of organs in males and females via binding to estrogen receptors. Upon receptor binding, E2 can trigger downstream signaling within the cell to regulate cell proliferation and gene transcription. Consequently, deregulated E2 can lead to the development of breast cancer and has been shown to determine the degree of breast cancer growth in 70% of all breast cancer cases[226]. Deregulated E2 can also cause increased cell migration and invasion[226]. As a result, inhibitors of E2 binding to estrogen receptors have been developed to treat breast cancers[226].

Oxygen gradients are also an important aspect of the TME. Due to rapid, uncontrolled proliferation, tumors have a tendency to outgrow their vasculature and grow faster than new blood vessels can form. This results in significant oxygen gradients leading to hypoxic pockets of cells. One of the major consequences of low oxygen levels is increased signaling via hypoxia-inducible factor (HIF), which results in the production of soluble signals that can affect epithelial-to-mesenchymal transition, angiogenesis, chemo- and radio-resistance, and metastasis, which are generally associated with poor clinical outcomes[217]. Furthermore, oxygen concentration has been shown to affect stem cell proportions within a tumor and select for radioresistant sub-populations[228] as well as the proliferative capacity of non-cancer cell types[219].

While soluble signaling can be evaluated in most bioengineered cancer models using techniques discussed below, a few models seek to isolate the effect of these soluble signals by preventing direct cell contacts between different cell types. For example, Regier et al. developed a compartmentalized system that allowed for tri-culture of breast cancer cells with stromal and immune components keeping each cell type physically separated while sharing the same culture medium. This allowed for direct analysis of paracrine signaling between co- and tri-cultures[229]. A similar model developed by Szot et al. was able to evaluate the role of paracrine signaling from breast cancer cells in causing angiogenesis of endothelial cells through culture of cancer cells and endothelial cells separated by an acellular type I collagen matrix[230]. It is important to note that models focused on paracrine signaling may yield unrepresentative results due to the absence of direct cell-cell signaling, though they are valuable for the evaluation of explicit paracrine effects.

Due to the pervasive effects of soluble signals in tumor progression and the dependence of soluble signaling on the culture model, soluble signals are often included in the analysis of cancer engineering models. However, the sub-micromolar concentrations of these signals within the TME complicates the design and capabilities of the models[221]. Typically, analysis occurs at the protein and/or RNA level using ELISA and RNA sequencing, respectively. Investigation of the effect of oxygen concentration requires cell culture in oxygen-controlled incubators and characterization of hypoxia using oxygen sensitive stains as well as analysis of protein and RNA level expression of hypoxia inducible factors. While methods of soluble signaling analysis are relatively uniform across labs, experimental variation is introduced by the wide variety of model systems implemented, which can affect the composition and concentration of soluble signals and thus differential cell behavior. This variation is caused by the incorporation of multiple cell types, culture dimensionality, mechanical stimuli, and immune cells in different cancer tissue engineering models.

The main limitation in ‘cancer-organ’ models that study soluble signaling is inherent in our inability to temporally recapitulate the in vivo microenvironment. The complexity of the environment often means that changing or omitting one component in the system can lead to a cascade of unrealistic effects within the tumor model. For example, it is known that MSCs can induce tumor cell proliferation and expression of VEGF[224], which in turn triggers angiogenesis and endothelial cell proliferation[231]. In turn, endothelial cells communicate with the tumor cells to drive progression[232,233]. In addition to the changes in tumor cells induced by MSC and endothelial cell signaling, the increased proliferation will also alter nutrient and oxygen gradients as well as mechanical forces present due to the increased tumor size, further affecting cancer cell behavior[77,79,147]. Therefore, any model that does not contain both endothelial cells and MSCs will result in only partial tumor cell changes which may lead to false conclusions. The same could be said for the omission of any aspect of the TME. Therefore, it is necessary to improve the complexity of our ‘cancer-organ’ model systems to preserve biologic relevance.

Finally, a point of concern for current models is that traditional cell culture is performed at oxygen concentrations around 20%, while physiologic oxygen levels in the TME are between 4 and 10%, meaning that most culture models are in hyperoxic conditions[217]. While some microfluidic devices modulate oxygen tension to physiologic levels[234236], other models of physiologic oxygen tension typically involve traditional cell culture in a multi-gas incubator with oxygen control[237,238]. These models operate under the assumption that the oxygen concentration that the cells sense is identical to that inside the incubator, while in reality it will depend on medium height and cell density[237,238]. The clear effect that oxygen concentrations have in the TME brings into question the reliability of in vitro cancer model experiments conducted at atmospheric oxygen levels as well as oxygen control incubators in the absence of cell level oxygen tension measurements. That said, hypoxic conditions overall have been well studied showing a clear influences on a variety of cellular mechanisms, such as promoting genetic instability, metastasis, invasiveness, and chemoresistance[239]. Thus, oxygen concentration levels should be of note and consideration for all ‘cancer-organ’ and TME investigations.

Complex and integrated ‘Cancer-Organ’ models

The lack of complexity in many in vitro models has led to poor efficacy in drug screening applications, leading to post-approval drug withdrawals. This is due to poor understanding of systemic drug toxicity and mechanism of action in a whole organ and precision oncology context[268]. Organ-on-a-chip models seek to remedy this problem through the integration of tissue engineering and microfluidics to develop more complex models that better recapitulate organ function and drug response. Specifically, these models aim to reproduce tissue microenvironments in terms of tissue level, multicellular organization, and tissue specific functions. Unlike organoid or tumoroid models, which rely on complex self-assembly of cells into organ like structures, organ-on-a-chip models provide more control over spatial confinement and allow for the connection of multiple organ systems. Recently, Zhang et al. developed an organ-on-a-chip model with built-in vasculature and a biodegradable scaffold made with poly(octamethylene maleate (anhydride) citrate). This chip allowed for parenchymal assembly on the matrix, which surrounded 3D, perfusable microchannels coated with endothelial cells. The permeability in the vessel walls allowed for enhanced intercellular communication as well as extravasation of monocytes and endothelial cells. Using this model, the authors created functionalized hepatic and cardiac tissues, which could be used to study the effect of drugs delivered through the vasculature[260]. Organ-on-a-chip devices have also been successfully developed to mimic the heart[260,261], liver[262,269], bone[263], kidney[270], lung[264], and gut[265] and have even been expanded to body-on-a-chip devices that integrate multiple organ systems into a single device[268]. Body-on-a-chip devices allow for inter-organ migration and communication as well as evaluation of systemic drug toxicity[268]. Culturing multiple “organs” on a single device can also serve as an ideal in vitro model of metastasis from a tumor to another organ. In fact, ‘patient-on-a-chip’ and ‘cancer/tumor-on-a-chip’ models are now being introduced to further individualize therapies.

Skardal et al. demonstrated this concept with the development of a metastasis-on-a-chip platform to study the metastasis of colon cancer cells co-cultured with epithelial intestinal cells in a ‘hyaluronic acid-polyethylene glycol diacrylate (PEGDA)-gelatin’ based hydrogel gut structure to a similar hydrogel containing HepG2 liver cells downstream in the microfluidic device[271]. This system also enabled mechanical tunability of the hydrogels via polymerization with linear, 4- arm, or 8-arm PEGDA to examine the effect of tissue stiffness on metastasis, showing increased metastatic ability in softer hydrogels[271].

Tumor-on-a-chip devices, like Skardal’s device, more accurately replicate the complex ‘organ-like’ microenvironment, leading to improved drug development and more accurate screening results compared to simpler in vitro models[272]. In fact, tumor-on-a-chip devices can overcome one of the main limitations persistent in most in vitro models: the lack of functional vasculature. The vasculature is an important component of in vitro models, as it provides oxygen and nutrients, delivers drugs and immune cells to the tumor, and serves as a channel for tumor cell migration and metastasis[272]. The endothelial cells that make up the vasculature also communicate dynamically with the tumor to direct tumor phenotype and angiogenesis via notch and vascular endothelial growth factor signaling, for example[232,233]. Tumor-on-a-chip devices demonstrate remarkable potential in regard to incorporation of realistic vascular components to a tumor model by creating vascularized channels within microfluidic devices.

One such device was fabricated using a polydimethylsiloxane (PDMS) microfluidic device with two outer channels (one arteriole and the other venule) connected by three ‘tissue chambers’. The tissue chambers were injected with endothelial and stromal cells with extracellular matrix. After 5–7 days of culture, the endothelial cells self-assembled and formed networks interconnected with the arteriole and venule channels directing flow within the lumen of the vascular network. Incorporation of human colorectal cancer cells into the tissue chambers resulted in the formation of tumor spheroids, which localized near the vessels and sometimes ensconced a region of the vessels recapitulating vasculature running through a tumor in vivo. Using this model, the authors were able to compare the effects of potential anti-cancer drugs, Pazopanib, Sorafenib, and Vincristine, on angiogenesis, maintenance of the vascular networks, and tumor spheroid growth[266].

Other advantages of tumor-on-a-chip devices include replication of key aspects of the TME like tumor-stroma interactions, tumor-ECM interactions, tumor-chemokine interactions[273], tumor-immune interactions, and complex processes like epithelial-to-mesenchymal transition and specific steps of metastasis like intravasation and extravasation[274]. Hao et al. developed a bone-on-a-chip device for co-culture of metastatic MDA-MB-231 breast cancer cells with mineralized collagenous bone tissue. Using this model, the authors studied breast cancer metastasis to bone and observed characteristics of breast cancer bone colonization that had previously only been seen in vivo, including rapid invasion of cancer cells into the apical layer of mineralized tissue, invadopodia that extended into distant matrix, cancer cells forming lines, and subsequent alignment of collagen parallel to the lines of cancer cells[263].

Finally, organ and tumor-on-a-chip devices can facilitate real-time analysis of experimental variables within the device[268]. This advantage was demonstrated in a liver-on-a-chip device that allowed for real-time analysis of metabolic function using a computer controlled microfluidic switchboard to measure glucose and lactate[262]. The ability of these devices to capture the complex multi-cellular, ‘organ-like’ environment of a tumor and neighboring organs combined with easy integration with analysis techniques make ‘on-a-chip’ devices promising platforms for the future of cancer research.

Mathematical modeling in the ‘Cancer-Organ’ models

The complexity of the TME makes perfect replication in vitro an unlikely feat; however, mathematical modeling can be used to help fill in gaps of knowledge left by incomplete models of complex ‘cancer-organs’. Specifically, mathematical modeling techniques can help fill in gaps in mechanistic understanding, indicate experiments that should be performed, and make personalized predictions of patient response to treatment[275,276]. This idea has led to the development of many mathematical models of cancer.

As we learn more about the complexity of the TME, the need for mathematical models becomes clearer, as accurate experimental models become more difficult to create. For example, the recent discovery of Tie2-expressing macrophages (TEMs), which influence tumor angiogenesis, vascular remodeling, and monocyte differentiation, added another macrophage phenotype into the system of macrophage interactions with the tumor[277]. To study macrophage interactions with the TME and the resulting tumor progression, a mathematical model was developed with M1, M2, and TEMs interacting with a metastatic lesion in a highly vascularized organ. The model included simulations of M1 release of nitric oxide, M2 release of general tumor growth factors, TEM secretion of angiopoietin-2 and IL-10, and evaluated their effects on tumor progression. This model revealed that TEM effects on tumor growth are irrelevant in the presence of M2 macrophages, suggesting that TEM targeting therapies would need to be administered in conjunction with M2-targeting therapies[277]. These findings provide insight into the weight of importance of TEMs vs M2 macrophages in tumor progression, which would be technically challenging to do via experimental means alone.

In another mathematical model, the authors sought to reconcile the development of resistance to epidermal growth factor receptor inhibitor in non-small cell lung cancer (NSCLC) with the changes in the TME. Specifically, they modeled the effect of glucose, oxygen, and drug concentrations on tumor evolution, with cell death and division rates dependent on the amount of available glucose, oxygen, and drug in a given region. In this model, cells with inherent drug resistance had different death and division rates, and drug sensitive cells had a small chance of developing resistance with each division. They found that compartments of their model with different oxygen, glucose, and drug concentrations had different predicted rebound times from drug treatment. They also found that the rebound following drug treatment varied depending on the mechanism of drug resistance and was independent of the number of resistant cells due to differential selective pressures in the different regions of the microenvironment[278]. This model helps to explain the role of tumor heterogeneity in the development of drug resistance and tumor recurrence and highlights the need to create more comprehensive experiments and mathematical models that take this heterogeneity into account.

Other more comprehensive models of the TME take into account multiple length scales to incorporate individual atomic events and cell interactions with discrete agent-based models, as well as bulk tumor growth, oxygen and nutrient diffusion, angiogenesis, and multi-organ processes, such as metastasis with continuous differential equation based models[240,279]. By merging multiple size and time scales, a more realistic depiction of the whole ‘cancer-organ’ can be obtained, resulting in higher model prediction accuracy[240]. While these hybrid models are more complex, advances in computation hardware has made them more feasible in recent years[240]. Despite the advantages that are inherent in the complexity of multiscale hybrid models, simpler models are still useful, as they are less computationally demanding and can still provide valuable biological insights.

The broad range of mathematical modeling complexity and techniques gives them almost unlimited potential in cancer research. Aside from the specific models discussed, many models have been created to examine almost any aspect of the TME including: tumor-immune interactions[277,280282], soluble signaling[277,282], tumor initiation and mutation rates[283], metastasis[284,285], angiogenesis[279,286,287], ligand binding events[282], development of tumor heterogeneity[276], proliferation and growth[288,289], dormancy[281,289], recurrence[278], extracellular matrix interactions[240,279], and drug resistance[278,290]. For a more detailed discussion of mathematical modeling techniques used to model the TME, readers are referred to several more comprehensive reviews focused on mathematical models of cancer[240,275,276]. Using these models to complement our experimental systems has the potential to add additional layers of complexity that would be difficult to study experimentally, thus helping to expand our understanding of the TME from a complex ‘organ-like’ perspective. Unfortunately, the specialized skill set required to implement these models serves as a barrier to the uniform adoption of mathematical models to supplement experimental research.

While cancer research has made many discoveries about the factors that influence tumor progression and developed new and improved treatments, we are still a long way from understanding the entire ‘cancer-organ’ system. More importantly, our lack of a complete understanding of tumor progression has slowed the development of novel treatments and potential cures. In the next part of this review, we will discuss the main challenges facing the field and propose solutions for improving physiologic and experimental reproducibility, as well as, the integration of biology and engineering.

Challenges and proposed solutions for developing physiologic ‘Cancer-Organ’ models

8.1 Physiological reproducibility

The inherent difficulty involved in building a functional ‘cancer-organ’ in the lab translates directly to in vitro tumor models by virtue of the similarities between the complex organization of an organ and the TME. Namely, it is immensely difficult to recapitulate the dimensionality, the mechanical environment, the cell-cell interactions, the soluble signals, and the immune compartment of the TME in a single model that is amenable to precise analysis. Each of these factors of the TME is integral in directing tumor progression as discussed above and thus should be considered in the design of experimental models. When many of these factors are not considered in a given model, there is an increased risk of producing unreliable results. Herein lies one of the main challenges facing cancer research: the lack of physiologically representative in vitro models due to the abundance of variables that exist within the complex, 3D, multicellular, organ-like TME. While a lot of progress has been made in terms of making more physiologic models, they are often still tailored to a specific research question and maintain reasonable simplicity via inclusion of only the necessary characteristics of the TME. For example, an investigation of the role of fibroblasts on cancer progression is likely to involve a 2- or 3-dimensional co-culture of fibroblasts with cancer cells. Such a co-culture neglects the other cell types within the TME, their cell-cell interactions and contributions to the soluble signaling milieu, and the mechanical stimuli. While it can be important to study one TME characteristic in isolation to understand its individual contributions to tumor biology, it is equally important to verify those findings in a more complex model with multiple TME characteristics. That said, a perfect replication of the organ-like TME of a tumor would be very technically challenging and would likely limit the reproducibility of experiments across labs. To overcome this limitation, cancer tissue-engineered models that present a slightly more comprehensive representation of the tumor while maintaining reasonable degrees of simplicity need to be developed to ensure that experimental results truly reflect the in vivo system.

It is also essential to develop cancer engineered models based on the different approaches required by solid tumors and liquid tumors to ensure that a model accurately reproduces in vivo physiology. Similar to solid tumor models, researchers have used cell lines and sophisticated animal models in addition to the more recent use of ex vivo models using hydrogels and scaffolds[291] in order to replicate the tumor microenvironment of liquid tumors. However, the characteristics of the microenvironment of liquid tumors, such as acute myeloid leukemia (AML), are significantly different from the microenvironment of solid tumors [292]. These differences need to be reflected in physiological models of liquid tumors compared to solid tumors to obtain realistic results. Therefore, selection of an appropriate personalized bioengineered model based on the type of cancer is fundamental for target discovery and precision oncology.

While the ideal model system would incorporate cellular diversity, dynamic mechanical forces, cell-cell and cell-ECM interactions, chemical gradients, and soluble signaling in a 3D microenvironment that is amenable to downstream analysis, we propose that researchers try to design their models to include at least three of the TME characteristics discussed in this review. Including additional TME characteristics into a culture model beyond the minimum necessary for the study will improve the robustness of the findings. If this is accomplished, our understanding of tumor progression will be greatly improved, and the identification of efficacious patient-specific treatments and treatment strategies will be accelerated. To best accomplish this, a multidisciplinary research approach involving collaborative efforts between engineers, oncologists, and immunologists is paramount to closely recapitulating the physiological TME and accelerating the race to cancers cures.

8.2 Experimental reproducibility

When discussing experimental reproducibility, special attention must be given to the vast number of variables that are modulated within experiments. As shown by the Sections above, there are endless attributes that can be modified independently and in conjunction with one another. Because of this, it is difficult to re-create a particular model used by another lab in order to directly verify a result. However, the variability between such models can be viewed as an advantage to the field. When two independent systems experimentally prod a specific characteristic of the tumor model and conclude the same or similar response tendency, i.e. reproducing each other’s findings, the result can be classified as robust. For example, serial 3D culture of ovarian cancer spheroids derived from cell lines as well as patient derived CSCs has been shown to increase resistance to cisplatin when formed via multiple different sphere formation methods including in 3D hanging drop plates[32], in low-attachment plates[293], and in traditional 2D culture with spontaneous spheroid budding[293] in separate labs. This conclusion being supported across multiple platforms is thus more reliable than a single platform alone. However, we find this is not always the case, and often similar experiments show contradictory results. This was the case when two separate labs knocked down ALDH1A1 with shRNA in ovarian cancer cell lines[294] or patient-derived ovarian cancer spheroids[295], showing that ALDH1A1 decreases[294] and increases[295] proliferation. This lack of reproducibility between labs draws attention to results which need to be investigated further. Much of this type of discrepancy can be attributed to unspecified conditions or slight differences in protocol that can have large effects on experimental outcome unbeknownst to the researchers.

To mitigate this conflict and improve overall reproducibility within the field, a more thorough and transparent documentation system should be practiced. Researchers should include detailed reports of their procedural characteristics and experimental failures as well. This would save not only time but also valuable funding resources and effort. Open source sharing of methodologies and recipes will help to establish consistency across experiments and facilitate best practices for studying a given microenvironment characteristic.

In peer-reviewed publications, reporting standards should be mandatory for experimental conditions, such as mechanical properties of the 3D culture system material (e.g. stiffness, porosity, permeability, and applicable rheometric values), biochemical properties of the model set up (e.g. cell adherence capabilities, oxygen diffusivity, and explicit medium compositions), and a thorough analysis of cellular profiles (e.g. genetic profile, cell subtype identification, plating densities, cell passage numbers, ratios of cell types). With this meticulous documentation, previously contradictory findings may be resolved and shed light on otherwise unidentified mechanisms within the TME. Additionally, the development of a standardized 3D model system that allows for fine control of each of the key TME characteristics (e.g. mechanical stimuli, co-culture, induced chemical gradients) would potentially eliminate dissimilar findings due to variation in experimental setups from different research groups.

8.3 Integration of biology and engineering

In the last 15 years, many transformative technologies have been developed to enhance the understanding of fundamental cancer biology and clinical translation. However, the integration of technologies and cancer biology has not been seamless, and as a result, the field has not made accelerated progress. Advances in molecular biology techniques allow for single cell downstream analysis; however, some techniques still require large number of cells like FACS, qRTPCR, and Western blot. Providing samples for such analysis without compromising the ability of an in vitro model to mimic the physiological TME is still technically challenging, cumbersome, and labor and time intensive. A key limitation to analysis techniques that require large cell numbers is the scarcity of patient samples, which most accurately represent in vivo phenotypes. To address the limitations imposed by assays that require large cell numbers to maintain accuracy, it is imperative to develop analysis methods that use low cell numbers. This will help to conserve rare cells like those available from patient samples and allow for multiple, high throughput, single cell assays, and effectively increase the information that can be obtained from a single patient sample. The effectiveness of this approach will improve with further development of minimalistic analysis techniques, such as drop sequencing, a single cell sequencing method, and automated liquid robotics enabled Western blot, which curtails sample volume needed for Western blotting.

There is also the issue of robustness of data acquired by different researchers for a single experiment, as user-to-user variability in technique leads to deviation in data. Success of the experiment depends on user experience, and most methods require learning new skills to ensure uniformity, robustness, and reproducibility. Often these skills become so specific to a subject that crossing the technological hurdles between biological and engineering fields becomes extremely difficult at a time when interdisciplinary research is essential for progress. This limitation highlights the need for detailed documentation of the methodologies discussed above. Often times, method sections will omit key experimental details that leave the reader extrapolating to fill in the gaps, which could lead to improper technique, unsuccessful experiments, and incongruent results, especially when interpreted by a less experienced researcher. As such, great care should be taken in drafting detailed methods that provide step-by-step instructions that researchers in different fields can follow. Such detailed methods will also help minimize experimental variation between researchers probing the same question. In addition, comprehensive reporting of model characterization will help to identify potential sources of variation between models from different research labs and thus aide in the accurate interpretation of results.

Furthermore, interdisciplinary efforts between dissimilar fields will help ensure a more seamless translation of model and analysis techniques between disciplines. For example, many physiologic models of cancers have been designed in engineering, physics, and chemistry labs with the intention of being widely adopted by the biology and clinical labs. However, the engineering parameters (such as fluid flow, ECM mechanics, etc.) might prove to be challenging to modulate for the biological and clinical labs. Evidence of this synergy between disciplines is already seen in collaborative grants, conferences, and publications that are being led by increasingly integrative teams in cancer biology, engineering, and oncology. One way to promote collaboration is through the promotion and development of more multidisciplinary conferences, which focus specifically on the inclusion of other fields to improve current models and techniques. These cross-disciplinary interactions are essential for the progress of the field and the improved design of experiments.

The power of interdisciplinary research in precision medicine is further evident in the development of advanced analytical techniques, which utilize concepts from physics, chemistry, and engineering to improve biological analysis of the more complex 3D in vitro tumor models present in research today. One such technique is live cell imaging performed with light sheet microscopy, which enables real-time live cell tracking and complete 3D imaging of the culture environment with minimal photobleaching, phototoxicity, and imaging time[296]. The information gleaned from such a technique may provide insight into key cell-cell interactions involved in tumor development and differentiation pathways and may help to explain experimental variation attributable to differences in 3D architecture[296], which would not otherwise be easily obtainable. Mass spectrometry-based proteomics is another advanced analytical technique that is becoming more widely implemented due to improvements in the mass spectrometry workflow and is now considered an irreplaceable molecular and cellular biology tool[297]. The powerful biological information provided by these advanced analysis techniques and other advanced imaging and ‘-omics’ analysis techniques in cancer research highlights the importance of interdisciplinary collaboration in enhancing progress in precision medicine[298].

Finally, increased development and utilization of mathematical models across subjects may serve as a common language through which biology, engineering, and chemistry can collaborate remotely, implement data to make predictions, and accelerate progress. The use of mathematical models in all fields is beneficial, as they can be created to analyze complex variables not easily studied experimentally, to determine which experiments are most promising, and to improve our understanding of biological mechanisms. These models could also decrease the use of scarce patient samples in experiments that can be determined in situ to not work and decrease the time needed to devise patient specific treatment plans. Integration will also be facilitated by the ease of sharing of codes used to generate mathematical models, effectively minimizing variance attributed to user differences and experience.

Conclusion

Overall, cancer can be considered a complex and interconnected organ system that colludes with its host in order to progress and maintain function. Our understanding of the ‘cancer-organ’ system relies on our ability to produce experimental models that accurately replicate critical aspects of the TME and provide reliable and meaningful results. In order to complete this task, cancer bioengineering models should consider the three dimensionalities of the tumor, the mechanical stimuli that continuously provoke response, the multicellular interactions innate to the environment, and the variety of sources that can provide signaling to a heterogeneous tumor. Understandably, each of these aspects encompasses numerous degrees of freedom, complicating the overarching picture. To remedy these challenges, we propose: 1) Enhancing physiological reproducibility through development of more comprehensive in vitro models, 2) Improving experimental reproducibility via reporting standards and sharing of negative results, 3) Sharing of knowledge and expertise across fields through collaboration, and 4) Improvement of analysis techniques to reduce technological hurdles. These improvements will facilitate the use of integrated platforms for studying tumor progression and organization, and developing the next generation of ‘cancer-organ’ models. Moreover, as a result of this work, we will gain significant understanding regarding the complex ways in which cancer cells interact with their surroundings. This has direct implications for both effective cancer prevention and individualized therapies and achieving better patient survival.

Acknowledgments

This work is supported primarily by DOD OCRP Early Career Investigator Award W81XWH-13-1-0134(GM), DOD Pilot award W81XWH-16-1-0426 (GM), DOD Investigator Initiated award W81XWH-17-OCRP-IIRA (GM), Rivkin Center for Ovarian Cancer (GM) and Michigan Ovarian Cancer Alliance (GM). Research reported in this publication was supported by the National Cancer Institute of the National Institutes of Health under award number P30CA046592. SR is supported by the National Institutes of Health under the award NIH/NIDCR T32DE00007057-41/42. CMN is supported by the National Science Foundation Graduate Research Fellowship under Grant No. 1256260. MEB is supported by the Department of Education Graduate Assistance in Areas of National Need (GAANN) Fellowship.

References

  1. 1. Hanahan D, Weinberg RA. Hallmarks of Cancer: The Next Generation. Cell. 2011;144: 646–674. pmid:21376230
  2. 2. Hanahan D, Weinberg RA. The Hallmarks of Cancer. Cell. 2000;100: 57–70. pmid:10647931
  3. 3. Heppner GH, Miller BE. Tumor heterogeneity: biological implications and therapeutic consequences. Cancer Metastasis Rev. 1983;2: 5–23. pmid:6616442
  4. 4. Holzel M, Bovier A, Tuting T. Plasticity of tumour and immune cells: a source of heterogeneity and a cause for therapy resistance? Nat Rev Cancer. 2013;13: 365–76. pmid:23535846
  5. 5. Egeblad M, Nakasone ES, Werb Z. Tumors as organs: complex tissues that interface with the entire organism. Dev Cell. 2010;18: 884–901. pmid:20627072
  6. 6. Loessner D, Stok KS, Lutolf MP, Hutmacher DW, Clements JA, Rizzi SC. Bioengineered 3D platform to explore cell–ECM interactions and drug resistance of epithelial ovarian cancer cells. Biomaterials. 2010;31: 8494–8506. pmid:20709389
  7. 7. Ip CKM, Li S-S, Tang MYH, Sy SKH, Ren Y, Shum HC, et al. Stemness and chemoresistance in epithelial ovarian carcinoma cells under shear stress. Sci Rep. 2016;6. pmid:27245437
  8. 8. Bougault C, Paumier A, Aubert-Foucher E, Mallein-Gerin F. Molecular analysis of chondrocytes cultured in agarose in response to dynamic compression. BMC Biotechnol. 2008;8: 71. pmid:18793425
  9. 9. Tse JM, Cheng G, Tyrrell JA, Wilcox-Adelman SA, Boucher Y, Jain RK, et al. Mechanical compression drives cancer cells toward invasive phenotype. Proc Natl Acad Sci. 2012;109: 911–916. pmid:22203958
  10. 10. Chin L, Xia Y, Discher DE, Janmey PA. Mechanotransduction in cancer. Curr Opin Chem Eng. 2016;11: 77–84. pmid:28344926
  11. 11. Takaishi K, Komohara Y, Tashiro H, Ohtake H, Nakagawa T, Katabuchi H, et al. Involvement of M2-polarized macrophages in the ascites from advanced epithelial ovarian carcinoma in tumor progression via Stat3 activation. Cancer Sci. 2010;101: 2128–2136. pmid:20860602
  12. 12. Yin M, Li X, Tan S, Zhou HJ, Ji W, Bellone S, et al. Tumor-associated macrophages drive spheroid formation during early transcoelomic metastasis of ovarian cancer. J Clin Invest. 2016;126: 4157–4173. pmid:27721235
  13. 13. Colvin EK. Tumor-Associated Macrophages Contribute to Tumor Progression in Ovarian Cancer. Front Oncol. 2014;4. pmid:24936477
  14. 14. Cassereau L, Miroshnikova YA, Ou G, Lakins J, Weaver VM. A 3D tension bioreactor platform to study the interplay between ECM stiffness and tumor phenotype. J Biotechnol. 2015;193: 66–69. pmid:25435379
  15. 15. Nargis NN, Aldredge RC, Guy RD. The influence of soluble fragments of extracellular matrix (ECM) on tumor growth and morphology. Math Biosci. 2018;296: 1–16. pmid:29208360
  16. 16. Alemany-Ribes M, Semino CE. Bioengineering 3D environments for cancer models. Adv Drug Deliv Rev. 2014;79–80: 40–49. pmid:24996134
  17. 17. Prina-Mello A, Jain N, Liu B, Kilpatrick JI, Tutty MA, Bell AP, et al. Culturing substrates influence the morphological, mechanical and biochemical features of lung adenocarcinoma cells cultured in 2D or 3D. Tissue Cell. 2018;50: 15–30. pmid:29429514
  18. 18. Le BD, Kang D, Yun S, Jeong YH, Jong-Young K, Yoon S, et al. Three-Dimensional Hepatocellular Carcinoma/Fibroblast Model on a Nanofibrous Membrane Mimics Tumor Cell Phenotypic Changes and Anticancer Drug Resistance. Nanomater Basel. 2018;8: 64. http://dx.doi.org/10.3390/nano8020064
  19. 19. Kinoshita T, Higuchi H, Ayano-Kabashima-Niibe , Sakai G, Hamamoto Y, Takaishi H, et al. Analysis of sensitivity and cell death pathways mediated by anti-cancer drugs using three-dimensional culture system. Int J Cancer Res. 2018;14: 1–12.
  20. 20. Ding Y, Liu W, Yu W, Lu S, Liu M, Kaplan DL, et al. Three-dimensional tissue culture model of human breast cancer for the evaluation of multidrug resistance. J Tissue Eng Regen Med. 2018;12: 1959–1971. pmid:30055109
  21. 21. Suo A, Xu W, Wang Y, Sun T, Ji L, Qian J. Dual-degradable and injectable hyaluronic acid hydrogel mimicking extracellular matrix for 3D culture of breast cancer MCF-7 cells. Carbohydr Polym. 2019;211: 336–348. pmid:30824098
  22. 22. DelNero P, Lane M, Verbridge SS, Kwee B, Kermani P, Hempstead B, et al. 3D culture broadly regulates tumor cell hypoxia response and angiogenesis via pro-inflammatory pathways. Biomaterials. 2015;55: 110–118. pmid:25934456
  23. 23. Edmondson R, Broglie JJ, Adcock AF, Yang L. Three-dimensional cell culture systems and their applications in drug discovery and cell-based biosensors. Assay Drug Dev Technol. 2014;12: 207–218. pmid:24831787
  24. 24. Imamura Y, Mukohara T, Shimono Y, Funakoshi Y, Chayahara N, Toyoda M, et al. Comparison of 2D- and 3D-culture models as drug-testing platforms in breast cancer. Oncol Rep. 2015;33: 1837-. pmid:25634491
  25. 25. Riedl A, Schlederer M, Pudelko K, Stadler M, Walter S, Unterleuthner D, et al. Comparison of cancer cells in 2D vs 3D culture reveals differences in AKT–mTOR–S6K signaling and drug responses. J Cell Sci. 2017;130: 203–218. pmid:27663511
  26. 26. Zimmermann M, Box C, Eccles SA. Two-Dimensional vs. Three-Dimensional In Vitro Tumor Migration and Invasion Assays. In: Moll J, Colombo R, editors. Target Identification and Validation in Drug Discovery: Methods and Protocols. Totowa, NJ: Humana Press; 2013. pp. 227–252. https://doi.org/10.1007/978-1-62703-311-4_15 pmid:23436416
  27. 27. Thoma CR, Zimmermann M, Agarkova I, Kelm JM, Krek W. 3D cell culture systems modeling tumor growth determinants in cancer target discovery. Adv Drug Deliv Rev. 2014;69–70: 29–41. pmid:24636868
  28. 28. Pampaloni F, Reynaud EG, Stelzer EHK. The third dimension bridges the gap between cell culture and live tissue. Nat Rev Mol Cell Biol. 2007;8: 839–845. pmid:17684528
  29. 29. Zhang M, Boughton P, Rose B, Lee CS, Hong AM. The Use of Porous Scaffold as a Tumor Model. Int J Biomater. 2013; pmid:24101930
  30. 30. Leslie K, Gao SP, Berishaj M, Podsypanina K, Ho H, Ivashkiv L, et al. Differential interleukin-6/Stat3 signaling as a function of cellular context mediates Ras-induced transformation. Breast Cancer Res BCR. 2010;12: R80. pmid:20929542
  31. 31. Pickl M, Ries CH. Comparison of 3D and 2D tumor models reveals enhanced HER2 activation in 3D associated with an increased response to trastuzumab. Oncogene. 2009;28: 461–468. pmid:18978815
  32. 32. Raghavan S, Mehta P, Horst EN, Ward MR, Rowley KR, Mehta G, et al. Comparative analysis of tumor spheroid generation techniques for differential in vitro drug toxicity. Oncotarget. 2016;7: 16948–16961. pmid:26918944
  33. 33. Raghavan S, Mehta P, Ward MW, Bregenzer ME, Fleck EMA, Tan L, et al. Personalized Medicine Based Approach to Model Patterns of Chemoresistance and Tumor Recurrence Using Ovarian Cancer Stem Cell Spheroids. Clin Cancer Res. 2017; pmid:28814433
  34. 34. Oliveira MB, Neto AI, Correia CR, Rial-Hermida MI, Alvarez-Lorenzo C, Mano JF. Superhydrophobic Chips for Cell Spheroids High-Throughput Generation and Drug Screening. ACS Appl Mater Interfaces. 2014;6: 9488–9495. pmid:24865973
  35. 35. Moshksayan K, Kashaninejad N, Warkiani ME, Lock JG, Moghadas H, Firoozabadi B, et al. Spheroids-on-a-chip: Recent advances and design considerations in microfluidic platforms for spheroid formation and culture. 2018; https://opus.lib.uts.edu.au/handle/10453/131013
  36. 36. Ho WY, Yeap SK, Ho CL, Rahim RA, Alitheen NB. Development of Multicellular Tumor Spheroid (MCTS) Culture from Breast Cancer Cell and a High Throughput Screening Method Using the MTT Assay. PLoS ONE. 2012;7. pmid:22970274
  37. 37. O’Brien FJ, Harley BA, Yannas IV, Gibson LJ. The effect of pore size on cell adhesion in collagen-GAG scaffolds. Biomaterials. 2005;26: 433–441. pmid:15275817
  38. 38. Liu J, Cheng F, Grénman H, Spoljaric S, Seppälä J, E. Eriksson J, et al. Development of nanocellulose scaffolds with tunable structures to support 3D cell culture. Carbohydr Polym. 2016;148: 259–271. pmid:27185139
  39. 39. Ulrich TA, Jain A, Tanner K, MacKay JL, Kumar S. Probing cellular mechanobiology in three-dimensional culture with collagen–agarose matrices. Biomaterials. 2010;31: 1875–1884. pmid:19926126
  40. 40. Afrimzon E, Botchkina G, Zurgil N, Shafran Y, Sobolev M, Moshkov S, et al. Hydrogel microstructure live-cell array for multiplexed analyses of cancer stem cells, tumor heterogeneity and differential drug response at single-element resolution. Lab Chip. 2016;16: 1047–1062. pmid:26907542
  41. 41. Lv D, Yu S, Ping Y, Wu H, Zhao X, Zhang H, et al. A three-dimensional collagen scaffold cell culture system for screening anti-glioma therapeutics. Oncotarget. 2016;7: 56904–56914. pmid:27486877
  42. 42. Wan L, Skoko J, Yu J, LeDuc PR, Neumann CA. Mimicking Embedded Vasculature Structure for 3D Cancer on a Chip Approaches through Micromilling. Sci Rep. 2017;7: 1–8.
  43. 43. Wang C, Li J, Sinha S, Peterson A, Grant GA, Yang F. Mimicking brain tumor-vasculature microanatomical architecture via co-culture of brain tumor and endothelial cells in 3D hydrogels. Biomaterials. 2019;202: 35–44. pmid:30836243
  44. 44. Martens P, Anseth KS. Characterization of hydrogels formed from acrylate modified poly(vinyl alcohol) macromers. Polymer. 2000;41: 7715–7722.
  45. 45. Hongisto V, Jernström S, Fey V, Mpindi J-P, Sahlberg KK, Kallioniemi O, et al. High-Throughput 3D Screening Reveals Differences in Drug Sensitivities between Culture Models of JIMT1 Breast Cancer Cells. PLOS ONE. 2013;8: e77232. pmid:24194875
  46. 46. Lee KY, Mooney DJ. Hydrogels for Tissue Engineering. Chem Rev. 2001;101: 1869–1880. pmid:11710233
  47. 47. Tibbitt MW, Anseth KS. Hydrogels as extracellular matrix mimics for 3D cell culture. Biotechnol Bioeng. 2009;103: 655–663. pmid:19472329
  48. 48. Kenny PA, Lee GY, Myers CA, Neve RM, Semeiks JR, Spellman PT, et al. The morphologies of breast cancer cell lines in three-dimensional assays correlate with their profiles of gene expression. Mol Oncol. 2007;1: 84–96. pmid:18516279
  49. 49. Worthington P, Pochan DJ, Langhans SA. Peptide Hydrogels—Versatile Matrices for 3D Cell Culture in Cancer Medicine. Front Oncol. 2015;5. pmid:25941663
  50. 50. Hughes CS, Postovit LM, Lajoie GA. Matrigel: A complex protein mixture required for optimal growth of cell culture. PROTEOMICS. 2010;10: 1886–1890. pmid:20162561
  51. 51. Szot CS, Buchanan CF, Freeman JW, Rylander MN. 3D in vitro bioengineered tumors based on collagen I hydrogels. Biomaterials. 2011;32: 7905–7912. pmid:21782234
  52. 52. Yue K, Trujillo-de Santiago G, Alvarez MM, Tamayol A, Annabi N, Khademhosseini A. Synthesis, properties, and biomedical applications of gelatin methacryloyl (GelMA) hydrogels. Biomaterials. 2015;73: 254–271. pmid:26414409
  53. 53. Kaemmerer E, Melchels FPW, Holzapfel BM, Meckel T, Hutmacher DW, Loessner D. Gelatine methacrylamide-based hydrogels: An alternative three-dimensional cancer cell culture system. Acta Biomater. 2014;10: 2551–2562. pmid:24590158
  54. 54. Gurski LA, Jha AK, Zhang C, Jia X, Farach-Carson MC. Hyaluronic acid-based hydrogels as 3D matrices for in vitro evaluation of chemotherapeutic drugs using poorly adherent prostate cancer cells. Biomaterials. 2009;30: 6076–6085. pmid:19695694
  55. 55. Ta HT, Dass CR, Dunstan DE. Injectable chitosan hydrogels for localised cancer therapy. J Controlled Release. 2008;126: 205–216. pmid:18258328
  56. 56. Hwang CM, Sant S, Masaeli M, Kachouie NN, Zamanian B, Lee S-H, et al. Fabrication of three-dimensional porous cell-laden hydrogel for tissue engineering. Biofabrication. 2010;2: 035003. pmid:20823504
  57. 57. Li L, Wang C, Huang Q, Xiao J, Zhang Q, Cheng Y. A degradable hydrogel formed by dendrimer-encapsulated platinum nanoparticles and oxidized dextran for repeated photothermal cancer therapy. J Mater Chem B. 2018;6: 2474–2480.
  58. 58. Fischbach C, Chen R, Matsumoto T, Schmelzle T, Brugge JS, Polverini PJ, et al. Engineering tumors with 3D scaffolds. Nat Methods. 2007;4: 855–860. pmid:17767164
  59. 59. Zhao W, Li J, Jin K, Liu W, Qiu X, Li C. Fabrication of functional PLGA-based electrospun scaffolds and their applications in biomedical engineering. Mater Sci Eng C. 2016;59: 1181–1194. pmid:26652474
  60. 60. Pan Z, Ding J. Poly(lactide-co-glycolide) porous scaffolds for tissue engineering and regenerative medicine. Interface Focus. 2012;2: 366–377. pmid:23741612
  61. 61. Fong ELS, Lamhamedi-Cherradi S-E, Burdett E, Ramamoorthy V, Lazar AJ, Kasper FK, et al. Modeling Ewing sarcoma tumors in vitro with 3D scaffolds. Proc Natl Acad Sci U S A. 2013;110: 6500–6505. pmid:23576741
  62. 62. Caicedo-Carvajal CE, Liu Q, Remache Y, Goy A, Suh KS. Cancer Tissue Engineering: A Novel 3D Polystyrene Scaffold for In Vitro Isolation and Amplification of Lymphoma Cancer Cells from Heterogeneous Cell Mixtures. J Tissue Eng. 2011;2011. pmid:22073378
  63. 63. Close DA, Camarco DP, Shan F, Kochanek SJ, Johnston PA. The Generation of Three-Dimensional Head and Neck Cancer Models for Drug Discovery in 384-Well Ultra-Low Attachment Microplates. In: Johnston PA, Trask OJ, editors. High Content Screening: A Powerful Approach to Systems Cell Biology and Phenotypic Drug Discovery. New York, NY: Springer New York; 2018. pp. 355–369. https://doi.org/10.1007/978-1-4939-7357-6_20 pmid:29082502
  64. 64. Mehta P, Novak C, Raghavan S, Ward M, Mehta G. Self-Renewal and CSCs In Vitro Enrichment: Growth as Floating Spheres. In: Papaccio G, Desiderio V, editors. Cancer Stem Cells: Methods and Protocols. New York, NY: Springer New York; 2018. pp. 61–75.
  65. 65. Raghavan S, Ward MR, Rowley KR, Wold RM, Takayama S, Buckanovich RJ, et al. Formation of stable small cell number three-dimensional ovarian cancer spheroids using hanging drop arrays for preclinical drug sensitivity assays. Gynecol Oncol. 2015;138: 181–189. pmid:25913133
  66. 66. Hyler AR, Baudoin NC, Brown MS, Stremler MA, Cimini D, Davalos RV, et al. Fluid shear stress impacts ovarian cancer cell viability, subcellular organization, and promotes genomic instability. PLOS ONE. 2018;13: e0194170. pmid:29566010
  67. 67. Jain RK, Martin JD, Stylianopoulos T. The Role of Mechanical Forces in Tumor Growth and Therapy. Annu Rev Biomed Eng. 2014;16: 321–346. pmid:25014786
  68. 68. DuFort CC, Paszek MJ, Weaver VM. Balancing forces: architectural control of mechanotransduction. Nat Rev Mol Cell Biol. 2011;12: 308–319. pmid:21508987
  69. 69. Novak C, Horst E, Mehta G. Review: Mechanotransduction in ovarian cancer: Shearing into the unknown. APL Bioeng. 2018;2: 031701.
  70. 70. Ao M, Brewer BM, Yang L, Franco Coronel OE, Hayward SW, Webb DJ, et al. Stretching Fibroblasts Remodels Fibronectin and Alters Cancer Cell Migration. Sci Rep. 2015;5: 8334. pmid:25660754
  71. 71. Chaudhuri O, Koshy ST, Branco da Cunha C, Shin J-W, Verbeke CS, Allison KH, et al. Extracellular matrix stiffness and composition jointly regulate the induction of malignant phenotypes in mammary epithelium. Nat Mater. 2014;13: 970–978. pmid:24930031
  72. 72. Fan R, Emery T, Zhang Y, Xia Y, Sun J, Wan J. Circulatory shear flow alters the viability and proliferation of circulating colon cancer cells. Sci Rep. 2016;6: 27073. pmid:27255403
  73. 73. Demou ZN. Gene Expression Profiles in 3D Tumor Analogs Indicate Compressive Strain Differentially Enhances Metastatic Potential. Ann Biomed Eng. 2010;38: 3509–3520. pmid:20559731
  74. 74. Aragona M, Panciera T, Manfrin A, Giulitti S, Michielin F, Elvassore N, et al. A Mechanical Checkpoint Controls Multicellular Growth through YAP/TAZ Regulation by Actin-Processing Factors. Cell. 2013;154: 1047–1059. pmid:23954413
  75. 75. Karjalainen HM, Sironen RK, Elo MA, Kaarniranta K, Takigawa M, Helminen HJ, et al. Gene expression profiles in chondrosarcoma cells subjected to cyclic stretching and hydrostatic pressure. A cDNA array study. Biorheology. 2003;40: 93–100. pmid:12454392
  76. 76. Yoo PS, Mulkeen AL, Dardik A, Cha CH. A Novel In Vitro Model of Lymphatic Metastasis from Colorectal Cancer. J Surg Res. 2007;143: 94–98. pmid:17640671
  77. 77. Voutouri C, Mpekris F, Papageorgis P, Odysseos AD, Stylianopoulos T. Role of Constitutive Behavior and Tumor-Host Mechanical Interactions in the State of Stress and Growth of Solid Tumors. Singh PK, editor. PLoS ONE. 2014;9: e104717. pmid:25111061
  78. 78. Li JF, Lowengrub J. The effects of cell compressibility, motility and contact inhibition on the growth of tumor cell clusters using the Cellular Potts Model. J Theor Biol. 2014;343: 79–91. pmid:24211749
  79. 79. Mpekris F, Angeli S, Pirentis AP, Stylianopoulos T. Stress-mediated progression of solid tumors: effect of mechanical stress on tissue oxygenation, cancer cell proliferation, and drug delivery. Biomech Model Mechanobiol. 2015;14: 1391–1402. pmid:25968141
  80. 80. Lynch ME, Chiou AE, Lee MJ, Marcott SC, Polamraju PV, Lee Y, et al. Three-Dimensional Mechanical Loading Modulates the Osteogenic Response of Mesenchymal Stem Cells to Tumor-Derived Soluble Signals. Tissue Eng Part A. 2016;22: 1006–1015. pmid:27401765
  81. 81. Marturano-Kruik A, Villasante A, Yaeger K, Ambati SR, Chramiec A, Raimondi MT, et al. Biomechanical regulation of drug sensitivity in an engineered model of human tumor. Biomaterials. 2018;150: 150–161. pmid:29040875
  82. 82. Kim BG, Gao M-Q, Kang S, Choi YP, Lee JH, Kim JE, et al. Mechanical compression induces VEGFA overexpression in breast cancer via DNMT3A-dependent miR-9 downregulation. Cell Death Dis. 2017;8: e2646. pmid:28252641
  83. 83. Srivastava N, Kay RR, Kabla AJ, Gardel M. Method to study cell migration under uniaxial compression. Mol Biol Cell. 2017;28: 809–816. pmid:28122819
  84. 84. Lam RHW, Weng S, Lu W, Fu J. Live-cell subcellular measurement of cell stiffness using a microengineered stretchable micropost array membrane. Integr Biol. 2012;4: 1289–1298. pmid:22935822
  85. 85. Paszek MJ, Zahir N, Johnson KR, Lakins JN, Rozenberg GI, Gefen A, et al. Tensional homeostasis and the malignant phenotype. Cancer Cell. 2005;8: 241–254. pmid:16169468
  86. 86. Nukuda A, Sasaki C, Ishihara S, Mizutani T, Nakamura K, Ayabe T, et al. Stiff substrates increase YAP-signaling-mediated matrix metalloproteinase-7 expression. Oncogenesis. 2015;4: e165. pmid:26344692
  87. 87. Liu Y, Lv J, Liang X, Yin X, Zhang L, Chen D, et al. Fibrin Stiffness Mediates Dormancy of Tumor-Repopulating Cells via a Cdc42-Driven Tet2 Epigenetic Program. Cancer Res. 2018;78: 3926–3937. pmid:29764867
  88. 88. Snook RD, Harvey TJ, Faria EC, Gardner P. Raman tweezers and their application to the study of singly trapped eukaryotic cells. Integr Biol. 2009;1: 43–52. pmid:20023790
  89. 89. Stowers RS, Allen SC, Suggs LJ. Dynamic phototuning of 3D hydrogel stiffness. Proc Natl Acad Sci. 2015;112: 1953–1958. pmid:25646417
  90. 90. Nam S, Lee J, Brownfield DG, Chaudhuri O. Viscoplasticity Enables Mechanical Remodeling of Matrix by Cells. Biophys J. 2016;111: 2296–2308. pmid:27851951
  91. 91. Wisdom KM, Adebowale K, Chang J, Lee JY, Nam S, Desai R, et al. Matrix mechanical plasticity regulates cancer cell migration through confining microenvironments. Nat Commun. 2018;9: 4144. pmid:30297715
  92. 92. Wullkopf L, West A-KV, Leijnse N, Cox TR, Madsen CD, Oddershede LB, et al. Cancer cells’ ability to mechanically adjust to extracellular matrix stiffness correlates with their invasive potential. Mol Biol Cell. 2018;29: 2378–2385. pmid:30091653
  93. 93. Huang Q, Hu X, He W, Zhao Y, Hao S, Wu Q, et al. Fluid shear stress and tumor metastasis. Am J Cancer Res. 2018;8: 763–777. pmid:29888101
  94. 94. Hou HW, Lee WC, Leong MC, Sonam S, Vedula SRK, Lim CT. Microfluidics for Applications in Cell Mechanics and Mechanobiology. Cell Mol Bioeng. 2011;4: 591–602.
  95. 95. Porquet N, Poirier A, Houle F, Pin A-L, Gout S, Tremblay P-L, et al. Survival advantages conferred to colon cancer cells by E-selectin-induced activation of the PI3K-NFκB survival axis downstream of Death receptor-3. BMC Cancer. 2011;11: 285. pmid:21722370
  96. 96. Lee HJ, Diaz MF, Price KM, Ozuna JA, Zhang S, Sevick-muraca EM, et al. Fluid shear stress activates YAP1 to promote cancer cell motility. Nat Commun Lond. 2017;8: 14122. http://dx.doi.org/10.1038/ncomms14122
  97. 97. Giavazzi R, Foppolo M, Dossi R, Remuzzi A. Rolling and adhesion of human tumor cells on vascular endothelium under physiological flow conditions. J Clin Invest. 1993;92: 3038–3044. pmid:7504697
  98. 98. Song H, David O, Clejan S, Giordano CL, Pappas-Lebeau H, Xu L, et al. Spatial Composition of Prostate Cancer Spheroids in Mixed and Static Cultures. Tissue Eng. 2004;10: 1266–1276. pmid:15363181
  99. 99. Whiteside TL. The tumor microenvironment and its role in promoting tumor growth. Oncogene. 2008;27: 5904–5912. pmid:18836471
  100. 100. Wang M, Zhao J, Zhang L, Wei F, Lian Y, Wu Y, et al. Role of tumor microenvironment in tumorigenesis. J Cancer. 2017;8: 761–773. pmid:28382138
  101. 101. Balkwill FR, Capasso M, Hagemann T. The tumor microenvironment at a glance. J Cell Sci. 2012;125: 5591–5596. pmid:23420197
  102. 102. Valkenburg KC, de Groot AE, Pienta KJ. Targeting the tumour stroma to improve cancer therapy. Nat Rev Clin Oncol. 2018;15: 366–381. pmid:29651130
  103. 103. Brücher BLDM, Jamall IS. Cell-Cell Communication in the Tumor Microenvironment, Carcinogenesis, and Anticancer Treatment. Cell Physiol Biochem. 2014;34: 213–243. pmid:25034869
  104. 104. Pang M-F, Nelson CM. Intercellular Communication, the Tumor Microenvironment, and Tumor Progression. In: Kandouz M, editor. Intercellular Communication in Cancer. Dordrecht: Springer Netherlands; 2015. pp. 343–362. https://doi.org/10.1007/978-94-017-7380-5_13
  105. 105. Stains JP, Civitelli R. Gap junctions regulate extracellular signal-regulated kinase signaling to affect gene transcription. Mol Biol Cell. 2005;16: 64–72. pmid:15525679
  106. 106. Yamasaki H. Gap Junctional Intercellular Communication and Carcinogenesis. In: Robards AW, Lucas WJ, Pitts JD, Jongsma HJ, Spray DC, editors. Parallels in Cell to Cell Junctions in Plants and Animals. Springer Berlin Heidelberg; 1990. pp. 115–127.
  107. 107. Eghbali B, Kessler JA, Reid LM, Roy C, Spray DC. Involvement of gap junctions in tumorigenesis: transfection of tumor cells with connexin 32 cDNA retards growth in vivo. Proc Natl Acad Sci U S A. 1991;88: 10701–10705. pmid:1660148
  108. 108. Kuphal S, Haass NK. Cell–Cell Contacts in Melanoma and the Tumor Microenvironment. In: Bosserhoff AK, editor. Melanoma Development: Molecular Biology, Genetics and Clinical Application. Cham: Springer International Publishing; 2017. pp. 227–269. https://doi.org/10.1007/978-3-319-41319-8_9
  109. 109. Runkle EA, Mu D. Tight Junction Proteins: From Barrier to Tumorigenesis. Cancer Lett. 2013;337: 41–48. pmid:23743355
  110. 110. Landy J, Ronde E, English N, Clark SK, Hart AL, Knight SC, et al. Tight junctions in inflammatory bowel diseases and inflammatory bowel disease associated colorectal cancer. World J Gastroenterol. 2016;22: 3117–3126. pmid:27003989
  111. 111. Martin TA, Jiang WG. Loss of tight junction barrier function and its role in cancer metastasis. Biochim Biophys Acta. 2009;1788: 872–891. pmid:19059202
  112. 112. Jiang WG, Sanders AJ, Katoh M, Ungefroren H, Gieseler F, Prince M, et al. Tissue invasion and metastasis: Molecular, biological and clinical perspectives. Semin Cancer Biol. 2015;35: S244–S275. pmid:25865774
  113. 113. Khan K, Hardy R, Haq A, Ogunbiyi O, Morton D, Chidgey M. Desmocollin switching in colorectal cancer. Br J Cancer. 2006;95: 1367–1370. pmid:17088906
  114. 114. Yang X, Wang J, Li W-P, Jin Z-J, Liu X-J. Desmocollin 3 mediates follicle stimulating hormone-induced ovarian epithelial cancer cell proliferation by activating the EGFR/Akt signaling pathway. Int J Clin Exp Pathol. 2015;8: 6716–6723. pmid:26261554
  115. 115. Zhou G, Yang L, Gray A, Srivastava AK, Li C, Zhang G, et al. The role of desmosomes in carcinogenesis. OncoTargets Ther. 2017;10: 4059–4063. pmid:28860814
  116. 116. Chidgey M, Dawson C. Desmosomes: a role in cancer? Br J Cancer. 2007;96: 1783–1787. pmid:17519903
  117. 117. Weiswald L-B, Bellet D, Dangles-Marie V. Spherical cancer models in tumor biology. Neoplasia N Y N. 2015;17: 1–15. pmid:25622895
  118. 118. Ham SL, Joshi R, Thakuri PS, Tavana H. Liquid-based three-dimensional tumor models for cancer research and drug discovery. Exp Biol Med Maywood NJ. 2016;241: 939–954. pmid:27072562
  119. 119. Weeber F, Ooft SN, Dijkstra KK, Voest EE. Tumor Organoids as a Pre-clinical Cancer Model for Drug Discovery. Cell Chem Biol. 2017;24: 1092–1100. pmid:28757181
  120. 120. Drost J, Clevers H. Organoids in cancer research. Nat Rev Cancer. 2018;18: 407. pmid:29692415
  121. 121. Mehta G, Hsiao AY, Ingram M, Luker GD, Takayama S. Opportunities and challenges for use of tumor spheroids as models to test drug delivery and efficacy. J Control Release Off J Control Release Soc. 2012; pmid:22613880
  122. 122. Sutherland RM, Inch WR, McCredie JA, Kruuv J. A multi-component radiation survival curve using an in vitro tumour model. Int J Radiat Biol Relat Stud Phys Chem Med. 1970;18: 491–495. pmid:5316564
  123. 123. Mayer B, Klement G, Kaneko M, Man S, Jothy S, Rak J, et al. Multicellular gastric cancer spheroids recapitulate growth pattern and differentiation phenotype of human gastric carcinomas. Gastroenterology. 2001;121: 839–852. pmid:11606498
  124. 124. Santini MT, Rainaldi G. Three-dimensional spheroid model in tumor biology. Pathobiol J Immunopathol Mol Cell Biol. 1999;67: 148–157. pmid:10394136
  125. 125. Haji-Karim M, Carlsson J. Proliferation and viability in cellular spheroids of human origin. Cancer Res. 1978;38: 1457–1464. pmid:25135
  126. 126. 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. pmid:20097238
  127. 127. Lawlor ER, Scheel C, Irving J, Sorensen PHB. Anchorage-independent multi-cellular spheroids as an in vitro model of growth signaling in Ewing tumors. Oncogene. 2002;21: 307–318. pmid:11803474
  128. 128. Ricci-Vitiani L, Lombardi DG, Pilozzi E, Biffoni M, Todaro M, Peschle C, et al. Identification and expansion of human colon-cancer-initiating cells. Nature. 2007;445: 111–115. pmid:17122771
  129. 129. Ponti D, Costa A, Zaffaroni N, Pratesi G, Petrangolini G, Coradini D, et al. Isolation and in vitro propagation of tumorigenic breast cancer cells with stem/progenitor cell properties. Cancer Res. 2005;65: 5506–5511. pmid:15994920
  130. 130. Dontu G, Abdallah WM, Foley JM, Jackson KW, Clarke MF, Kawamura MJ, et al. In vitro propagation and transcriptional profiling of human mammary stem/progenitor cells. Genes Dev. 2003;17: 1253–1270. pmid:12756227
  131. 131. Bjerkvig R, Tønnesen A, Laerum OD, Backlund EO. Multicellular tumor spheroids from human gliomas maintained in organ culture. J Neurosurg. 1990;72: 463–475. pmid:2406382
  132. 132. Tonn JC, Ott MM, Meixensberger J, Paulus W, Roosen K. Progesterone receptors are detectable in tumor fragment spheroids of meningiomas in vitro. Anticancer Res. 1994;14: 2453–2456. pmid:7872666
  133. 133. Rajcevic U, Knol JC, Piersma S, Bougnaud S, Fack F, Sundlisaeter E, et al. Colorectal cancer derived organotypic spheroids maintain essential tissue characteristics but adapt their metabolism in culture. Proteome Sci. 2014;12: 39. pmid:25075203
  134. 134. Spheroid Culture in Cancer Research (1991). In: Taylor & Francis [Internet]. 20 Dec 2018 [cited 20 Dec 2018]. https://www.taylorfrancis.com/books/e/9781351357630
  135. 135. Xu H, Lyu X, Yi M, Zhao W, Song Y, Wu K. Organoid technology and applications in cancer research. J Hematol OncolJ Hematol Oncol. 2018;11. pmid:30219074
  136. 136. Nagle PW, Plukker JTM, Muijs CT, van Luijk P, Coppes RP. Patient-derived tumor organoids for prediction of cancer treatment response. Semin Cancer Biol. 2018;53: 258–264. pmid:29966678
  137. 137. Sachs N, de Ligt J, Kopper O, Gogola E, Bounova G, Weeber F, et al. A Living Biobank of Breast Cancer Organoids Captures Disease Heterogeneity. Cell. 2018;172: 373–386.e10. pmid:29224780
  138. 138. Driehuis E, Clevers H. CRISPR/Cas 9 genome editing and its applications in organoids. Am J Physiol Gastrointest Liver Physiol. 2017;312: G257–G265. pmid:28126704
  139. 139. Zhang Z, Zhang Y, Gao F, Han S, Cheah KS, Tse H-F, et al. CRISPR/Cas9 Genome-Editing System in Human Stem Cells: Current Status and Future Prospects. Mol Ther Nucleic Acids. 2017;9: 230–241. pmid:29246302
  140. 140. Cruz NM, Song X, Czerniecki SM, Gulieva RE, Churchill AJ, Kim YK, et al. Organoid cystogenesis reveals a critical role of microenvironment in human polycystic kidney disease. Nat Mater. 2017;16: 1112–1119. pmid:28967916
  141. 141. Takasato M, Er PX, Chiu HS, Maier B, Baillie GJ, Ferguson C, et al. Kidney organoids from human iPS cells contain multiple lineages and model human nephrogenesis. Nature. 2015;526: 564–568. pmid:26444236
  142. 142. Barkauskas CE, Chung M-I, Fioret B, Gao X, Katsura H, Hogan BLM. Lung organoids: current uses and future promise. Development. 2017;144: 986–997. pmid:28292845
  143. 143. Huch M, Gehart H, van Boxtel R, Hamer K, Blokzijl F, Verstegen MMA, et al. Long-Term Culture of Genome-Stable Bipotent Stem Cells from Adult Human Liver. Cell. 2015;160: 299–312. pmid:25533785
  144. 144. Broutier L, Andersson-Rolf A, Hindley CJ, Boj SF, Clevers H, Koo B-K, et al. Culture and establishment of self-renewing human and mouse adult liver and pancreas 3D organoids and their genetic manipulation. Nat Protoc. 2016;11: 1724–1743. pmid:27560176
  145. 145. Lancaster MA, Renner M, Martin C-A, Wenzel D, Bicknell LS, Hurles ME, et al. Cerebral organoids model human brain development and microcephaly. Nature. 2013;501: 373–379. pmid:23995685
  146. 146. Sato T, Stange DE, Ferrante M, Vries RGJ, van Es JH, van den Brink S, et al. Long-term Expansion of Epithelial Organoids From Human Colon, Adenoma, Adenocarcinoma, and Barrett’s Epithelium. Gastroenterology. 2011;141: 1762–1772. pmid:21889923
  147. 147. Li X, Nadauld L, Ootani A, Corney DC, Pai RK, Gevaert O, et al. Oncogenic transformation of diverse gastrointestinal tissues in primary organoid culture. Nat Med. 2014;20: 769–777. pmid:24859528
  148. 148. Barker N, van Es JH, Kuipers J, Kujala P, van den Born M, Cozijnsen M, et al. Identification of stem cells in small intestine and colon by marker gene Lgr5. Nature. 2007;449: 1003–1007. pmid:17934449
  149. 149. Bijsmans ITGW, Milona A, Ijssennagger N, Willemsen ECL, Ramos Pittol JM, Jonker JW, et al. Characterization of stem cell-derived liver and intestinal organoids as a model system to study nuclear receptor biology. Biochim Biophys Acta BBA—Mol Basis Dis. 2017;1863: 687–700. pmid:27956139
  150. 150. Gao D, Vela I, Sboner A, Iaquinta PJ, Karthaus WR, Gopalan A, et al. Organoid Cultures Derived from Patients with Advanced Prostate Cancer. Cell. 2014;159: 176–187. pmid:25201530
  151. 151. Hubert CG, Rivera M, Spangler LC, Wu Q, Mack SC, Prager BC, et al. A Three-Dimensional Organoid Culture System Derived from Human Glioblastomas Recapitulates the Hypoxic Gradients and Cancer Stem Cell Heterogeneity of Tumors Found In Vivo. Cancer Res. 2016;76: 2465–2477. pmid:26896279
  152. 152. Huang L, Holtzinger A, Jagan I, BeGora M, Lohse I, Ngai N, et al. Ductal pancreatic cancer modeling and drug screening using human pluripotent stem cell—and patient-derived tumor organoids. Nat Med. 2015;21: 1364–1371. pmid:26501191
  153. 153. Boj SF, Hwang C-I, Baker LA, Chio IIC, Engle DD, Corbo V, et al. Organoid Models of Human and Mouse Ductal Pancreatic Cancer. Cell. 2015;160: 324–338. pmid:25557080
  154. 154. Walsh AJ, Castellanos JA, Nagathihalli NS, Merchant NB, Skala MC. Optical Imaging of Drug-Induced Metabolism Changes in Murine and Human Pancreatic Cancer Organoids Reveals Heterogeneous Drug Response: Pancreas. 2016;45: 863–869. pmid:26495796
  155. 155. van de Wetering M, Francies HE, Francis JM, Bounova G, Iorio F, Pronk A, et al. Prospective Derivation of a Living Organoid Biobank of Colorectal Cancer Patients. Cell. 2015;161: 933–945. pmid:25957691
  156. 156. Serban MA, Prestwich GD. Modular Extracellular Matrices: Solutions for the Puzzle. Methods San Diego Calif. 2008;45: 93–98. pmid:18442709
  157. 157. Trisno SL, Philo KED, McCracken KW, Catá EM, Ruiz-Torres S, Rankin SA, et al. Esophageal Organoids from Human Pluripotent Stem Cells Delineate Sox2 Functions during Esophageal Specification. Cell Stem Cell. 2018;23: 501–515.e7. pmid:30244869
  158. 158. Raghavan S, Mehta P, Horst EN, Ward MR, Rowley KR, Mehta G, et al. Comparative analysis of tumor spheroid generation techniques for differential in vitro drug toxicity. Oncotarget. 2016;7: 16948–16961. pmid:26918944
  159. 159. Hoarau-Véchot J, Rafii A, Touboul C, Pasquier J. Halfway between 2D and Animal Models: Are 3D Cultures the Ideal Tool to Study Cancer-Microenvironment Interactions? Int J Mol Sci. 2018;19. pmid:29346265
  160. 160. Eder T, Weber A, Neuwirt H, Grünbacher G, Ploner C, Klocker H, et al. Cancer-Associated Fibroblasts Modify the Response of Prostate Cancer Cells to Androgen and Anti-Androgens in Three-Dimensional Spheroid Culture. Int J Mol Sci. 2016;17. pmid:27598125
  161. 161. Asghar W, El Assal R, Shafiee H, Pitteri S, Paulmurugan R, Demirci U. Engineering cancer microenvironments for in vitro 3-D tumor models. Mater Today Kidlington Engl. 2015;18: 539–553. pmid:28458612
  162. 162. Gajewski TF, Schreiber H, Fu YX. Innate and adaptive immune cells in the tumor microenvironment. Nat Immunol. 2013;14: 1014–22. pmid:24048123
  163. 163. Galon J, Costes A, Sanchez-Cabo F, Kirilovsky A, Mlecnik B, Lagorce-Pages C, et al. Type, density, and location of immune cells within human colorectal tumors predict clinical outcome. Science. 2006;313: 1960–4. pmid:17008531
  164. 164. Kerkar SP, Restifo NP. Cellular constituents of immune escape within the tumor microenvironment. Cancer Res. 2012;72: 3125–30. pmid:22721837
  165. 165. Azimi F, Scolyer RA, Rumcheva P, Moncrieff M, Murali R, McCarthy SW, et al. Tumor-infiltrating lymphocyte grade is an independent predictor of sentinel lymph node status and survival in patients with cutaneous melanoma. J Clin Oncol. 2012;30: 2678–83. pmid:22711850
  166. 166. Mahmoud SM, Paish EC, Powe DG, Macmillan RD, Grainge MJ, Lee AH, et al. Tumor-infiltrating CD8+ lymphocytes predict clinical outcome in breast cancer. J Clin Oncol. 2011;29: 1949–55. pmid:21483002
  167. 167. Zhang L, Conejo-Garcia JR, Katsaros D, Gimotty PA, Massobrio M, Regnani G, et al. Intratumoral T cells, recurrence, and survival in epithelial ovarian cancer. N Engl J Med. 2003;348: 203–13. pmid:12529460
  168. 168. Rusakiewicz S, Semeraro M, Sarabi M, Desbois M, Locher C, Mendez R, et al. Immune infiltrates are prognostic factors in localized gastrointestinal stromal tumors. Cancer Res. 2013;73: 3499–510. pmid:23592754
  169. 169. Huber S, Hoffmann R, Muskens F, Voehringer D. Alternatively activated macrophages inhibit T-cell proliferation by Stat6-dependent expression of PD-L2. Blood. 2010;116: 3311–20. pmid:20625006
  170. 170. Biswas SK, Mantovani A. Macrophage plasticity and interaction with lymphocyte subsets: cancer as a paradigm. Nat Immunol. 2010;11: 889–96. pmid:20856220
  171. 171. Mantovani A, Sozzani S, Locati M, Allavena P, Sica A. Macrophage polarization: tumor-associated macrophages as a paradigm for polarized M2 mononuclear phagocytes. Trends Immunol. 2002;23: 549–55. pmid:12401408
  172. 172. Noy R, Pollard JW. Tumor-associated macrophages: from mechanisms to therapy. Immunity. 2014;41: 49–61. pmid:25035953
  173. 173. Yang L, Zhang Y. Tumor-associated macrophages: from basic research to clinical application. J Hematol Oncol. 2017;10: 58. pmid:28241846
  174. 174. Old LJ, Boyse EA. Immunology of Experimental Tumors. Annu Rev Med. 1964;15: 167–86. pmid:14139934
  175. 175. Penn I, Starzl TE. Malignant tumors arising de novo in immunosuppressed organ transplant recipients. Transplantation. 1972;14: 407–17. pmid:4345337
  176. 176. Schreiber RD, Old LJ, Smyth MJ. Cancer immunoediting: integrating immunity’s roles in cancer suppression and promotion. Science. 2011;331: 1565–70. pmid:21436444
  177. 177. Swann JB, Vesely MD, Silva A, Sharkey J, Akira S, Schreiber RD, et al. Demonstration of inflammation-induced cancer and cancer immunoediting during primary tumorigenesis. Proc Natl Acad Sci U A. 2008;105: 652–6. pmid:18178624
  178. 178. Dunn GP, Bruce AT, Ikeda H, Old LJ, Schreiber RD. Cancer immunoediting: from immunosurveillance to tumor escape. Nat Immunol. 2002;3: 991–8. pmid:12407406
  179. 179. Mittal D, Gubin MM, Schreiber RD, Smyth MJ. New insights into cancer immunoediting and its three component phases—elimination, equilibrium and escape. Curr Opin Immunol. 2014;27: 16–25. pmid:24531241
  180. 180. Sun C, Lan P, Han Q, Huang M, Zhang Z, Xu G, et al. Oncofetal gene SALL4 reactivation by hepatitis B virus counteracts miR-200c in PD-L1-induced T cell exhaustion. Nat Commun. 2018;9: 1241. pmid:29593314
  181. 181. Diehl R, Ferrara F, Muller C, Dreyer AY, McLeod DD, Fricke S, et al. Immunosuppression for in vivo research: state-of-the-art protocols and experimental approaches. Cell Mol Immunol. 2017;14: 146–179. pmid:27721455
  182. 182. Chaganty BKR, Qiu S, Gest A, Lu Y, Ivan C, Calin GA, et al. Trastuzumab upregulates PD-L1 as a potential mechanism of trastuzumab resistance through engagement of immune effector cells and stimulation of IFNgamma secretion. Cancer Lett. 2018;430: 47–56. pmid:29746929
  183. 183. Trumpi K, Frenkel N, Peters T, Korthagen NM, Jongen JMJ, Raats D, et al. Macrophages induce “budding” in aggressive human colon cancer subtypes by protease-mediated disruption of tight junctions. Oncotarget. 2018;9: 19490–19507. pmid:29731961
  184. 184. Rijal G, Li W. Native-mimicking in vitro microenvironment: an elusive and seductive future for tumor modeling and tissue engineering. J Biol Eng. 2018;12: 20. pmid:30220913
  185. 185. Holzapfel BM, Wagner F, Thibaudeau L, Levesque JP, Hutmacher DW. Concise review: humanized models of tumor immunology in the 21st century: convergence of cancer research and tissue engineering. Stem Cells. 2015;33: 1696–704. pmid:25694194
  186. 186. Kumar V, Varghese S. Ex Vivo Tumor-on-a-Chip Platforms to Study Intercellular Interactions within the Tumor Microenvironment. Adv Heal Mater. 2018; e1801198. pmid:30516355
  187. 187. Verhoeckx K, Cotter P, European Cooperation in the Field of Scientific and Technical Research (Organization). The impact of food bioactives on gut health: in vitro and ex vivo models. 2015.
  188. 188. Rizvi NA, Hellmann MD, Snyder A, Kvistborg P, Makarov V, Havel JJ, et al. Mutational landscape determines sensitivity to PD-1 blockade in non–small cell lung cancer. Science. 2015; aaa1348.
  189. 189. Nagasaki M, Yasuda J, Katsuoka F, Nariai N, Kojima K, Kawai Y, et al. Rare variant discovery by deep whole-genome sequencing of 1,070 Japanese individuals. Nat Commun. 2015;6: 8018. pmid:26292667
  190. 190. Gros A, Parkhurst MR, Tran E, Pasetto A, Robbins PF, Ilyas S, et al. Prospective identification of neoantigen-specific lymphocytes in the peripheral blood of melanoma patients. Nat Med. 2016;22: 433–8. pmid:26901407
  191. 191. Aravanis AM, Lee M, Klausner RD. Next-Generation Sequencing of Circulating Tumor DNA for Early Cancer Detection. Cell. 2017;168: 571–574. pmid:28187279
  192. 192. Hackl H, Charoentong P, Finotello F, Trajanoski Z. Computational genomics tools for dissecting tumour-immune cell interactions. Nat Rev Genet. 2016;17: 441–58. pmid:27376489
  193. 193. Herter S, Morra L, Schlenker R, Sulcova J, Fahrni L, Waldhauer I, et al. A novel three-dimensional heterotypic spheroid model for the assessment of the activity of cancer immunotherapy agents. Cancer Immunol Immunother. 2017;66: 129–140. pmid:27858101
  194. 194. Finnberg NK, Gokare P, Lev A, Grivennikov SI, MacFarlane AW th, Campbell KS, et al. Application of 3D tumoroid systems to define immune and cytotoxic therapeutic responses based on tumoroid and tissue slice culture molecular signatures. Oncotarget. 2017;8: 66747–66757. pmid:28977993
  195. 195. Ghosh S, Rosenthal R, Zajac P, Weber WP, Oertli D, Heberer M, et al. Culture of melanoma cells in 3-dimensional architectures results in impaired immunorecognition by cytotoxic T lymphocytes specific for Melan-A/MART-1 tumor-associated antigen. Ann Surg. 2005;242: 851–7, discussion 858. pmid:16327495
  196. 196. Linde N, Gutschalk CM, Hoffmann C, Yilmaz D, Mueller MM. Integrating macrophages into organotypic co-cultures: a 3D in vitro model to study tumor-associated macrophages. PLoS One. 2012;7: e40058. pmid:22792213
  197. 197. Hutmacher DW, Horch RE, Loessner D, Rizzi S, Sieh S, Reichert JC, et al. Translating tissue engineering technology platforms into cancer research. J Cell Mol Med. 2009;13: 1417–27. pmid:19627398
  198. 198. Jeanbart L, Swartz MA. Engineering opportunities in cancer immunotherapy. Proc Natl Acad Sci U A. 2015;112: 14467–72. pmid:26598681
  199. 199. Irvine DJ, Stachowiak AN, Hori Y. Lymphoid tissue engineering: invoking lymphoid tissue neogenesis in immunotherapy and models of immunity. Semin Immunol. 2008;20: 137–46. pmid:18035552
  200. 200. Grupp SA, Kalos M, Barrett D, Aplenc R, Porter DL, Rheingold SR, et al. Chimeric Antigen Receptor–Modified T Cells for Acute Lymphoid Leukemia. N Engl J Med. 2013;368: 1509–1518. pmid:23527958
  201. 201. Newick K, Moon E, Albelda SM. Chimeric antigen receptor T-cell therapy for solid tumors. Mol Ther Oncolytics. 2016;3: 16006. pmid:27162934
  202. 202. Brown CE, Alizadeh D, Starr R, Weng L, Wagner JR, Naranjo A, et al. Regression of Glioblastoma after Chimeric Antigen Receptor T-Cell Therapy. N Engl J Med. 2016;375: 2561–2569. pmid:28029927
  203. 203. Neelapu SS, Tummala S, Kebriaei P, Wierda W, Gutierrez C, Locke FL, et al. Chimeric antigen receptor T-cell therapy—assessment and management of toxicities. Nat Rev Clin Oncol. 2018;15: 47–62. pmid:28925994
  204. 204. Maus MV, Haas AR, Beatty GL, Albelda SM, Levine BL, Liu X, et al. T cells expressing chimeric antigen receptors can cause anaphylaxis in humans. Cancer Immunol Res. 2013;1: 26–31. pmid:24777247
  205. 205. Grigor EJM, Fergusson D, Kekre N, Montroy J, Atkins H, Seftel M, et al. Risks and Benefits of Chimeric Antigen Receptor T-Cell (CAR-T) Therapy in Cancer: A Systematic Review and Meta-Analysis. Transfus Med Rev. 2019; pmid:30948292
  206. 206. Mata M, Gottschalk S. Engineering for Success: Approaches to Improve Chimeric Antigen Receptor T Cell Therapy for Solid Tumors. Drugs. 2019; pmid:30796733
  207. 207. Yu S, Li A, Liu Q, Li T, Yuan X, Han X, et al. Chimeric antigen receptor T cells: a novel therapy for solid tumors. J Hematol OncolJ Hematol Oncol. 2017;10: 78. pmid:28356156
  208. 208. Buonaguro L, Pulendran B. Immunogenomics and systems biology of vaccines. Immunol Rev. 2011;239: 197–208. pmid:21198673
  209. 209. Sen P, Kemppainen E, Orešič M. Perspectives on Systems Modeling of Human Peripheral Blood Mononuclear Cells. Front Mol Biosci. 2018;4. pmid:29376056
  210. 210. Germain RN, Meier-Schellersheim M, Nita-Lazar A, Fraser IDC. Systems Biology in Immunology—A Computational Modeling Perspective. Annu Rev Immunol. 2011;29: 527–585. pmid:21219182
  211. 211. Zhao Y, Tian B, Edeh CB, Brasier AR. Quantitation of the dynamic profiles of the innate immune response using multiplex selected reaction monitoring-mass spectrometry. Mol Cell Proteomics. 2013;12: 1513–29. pmid:23418394
  212. 212. Varadarajan N, Julg B, Yamanaka YJ, Chen H, Ogunniyi AO, McAndrew E, et al. A high-throughput single-cell analysis of human CD8(+) T cell functions reveals discordance for cytokine secretion and cytolysis. J Clin Invest. 2011;121: 4322–31. pmid:21965332
  213. 213. Jacobs JF, Idema AJ, Bol KF, Grotenhuis JA, de Vries IJ, Wesseling P, et al. Prognostic significance and mechanism of Treg infiltration in human brain tumors. J Neuroimmunol. 2010;225: 195–9. pmid:20537408
  214. 214. Mitsui J, Nishikawa H, Muraoka D, Wang L, Noguchi T, Sato E, et al. Two distinct mechanisms of augmented antitumor activity by modulation of immunostimulatory/inhibitory signals. Clin Cancer Res. 2010;16: 2781–91. pmid:20460483
  215. 215. Sutmuller RP, van Duivenvoorde LM, van Elsas A, Schumacher TN, Wildenberg ME, Allison JP, et al. Synergism of cytotoxic T lymphocyte-associated antigen 4 blockade and depletion of CD25(+) regulatory T cells in antitumor therapy reveals alternative pathways for suppression of autoreactive cytotoxic T lymphocyte responses. J Exp Med. 2001;194: 823–32. pmid:11560997
  216. 216. Klebanoff CA, Khong HT, Antony PA, Palmer DC, Restifo NP. Sinks, suppressors and antigen presenters: how lymphodepletion enhances T cell-mediated tumor immunotherapy. Trends Immunol. 2005;26: 111–7. pmid:15668127
  217. 217. Muz B, de la Puente P, Azab F, Azab AK. The role of hypoxia in cancer progression, angiogenesis, metastasis, and resistance to therapy. Hypoxia. 2015; 83. pmid:27774485
  218. 218. AbdelFattah HS, Ibrahim MT, Nasr MM, Nasr Amin SN. Cell Signaling in Cancer Microenvironment. Int J Adv Biomed. 2017;2: 47–51.
  219. 219. Bhome R, Bullock MD, Al Saihati HA, Goh RW, Primrose JN, Sayan AE, et al. A top-down view of the tumor microenvironment: structure, cells and signaling. Front Cell Dev Biol. 2015;3. pmid:26075202
  220. 220. Gales D, Clark C, Manne U, Samuel T. The Chemokine CXCL8 in Carcinogenesis and Drug Response. ISRN Oncol. 2013;2013: 1–8. pmid:24224100
  221. 221. Kang D-K, Lu J, Kang D-K, Lu J, Zhang W, Chang E, et al. Microfluidic devices for stem cell analysis. Microfluidic Devices for Biomedical Applications. Elsevier; 2013. pp. 388–441.
  222. 222. Deng Y, Yang Y, Yao B, Ma L, Wu Q, Yang Z, et al. Paracrine signaling by VEGF-C promotes non-small cell lung cancer cell metastasis via recruitment of tumor-associated macrophages. Exp Cell Res. 2018;364: 208–216. pmid:29427623
  223. 223. Kuzet S-E, Gaggioli C. Fibroblast activation in cancer: when seed fertilizes soil. Cell Tissue Res. 2016;365: 607–619. pmid:27474009
  224. 224. Zhu W, Huang L, Li Y, Qian H, Shan X, Yan Y, et al. Mesenchymal stem cell-secreted soluble signaling molecules potentiate tumor growth. Cell Cycle. 2011;10: 3198–3207. pmid:21900753
  225. 225. Rolny C, Mazzone M, Tugues S, Laoui D, Johansson I, Coulon C, et al. HRG Inhibits Tumor Growth and Metastasis by Inducing Macrophage Polarization and Vessel Normalization through Downregulation of PlGF. Cancer Cell. 2011;19: 31–44. pmid:21215706
  226. 226. Busonero C, Leone S, Klemm C, Acconcia F. A functional drug re-purposing screening identifies carfilzomib as a drug preventing 17β-estradiol: ERα signaling and cell proliferation in breast cancer cells. Mol Cell Endocrinol. 2018;460: 229–237. pmid:28760601
  227. 227. Park J, Euhus DM, Scherer PE. Paracrine and Endocrine Effects of Adipose Tissue on Cancer Development and Progression. Endocr Rev. 2011;32: 550–570. pmid:21642230
  228. 228. Kim H, Lin Q, Yun Z. The hypoxic tumor microenvironment in vivo selects tumor cells with increased survival against genotoxic stresses. Cancer Lett. 2018;431: 142–149. pmid:29859297
  229. 229. Regier MC, Maccoux LJ, Weinberger EM, Regehr KJ, Berry SM, Beebe DJ, et al. Transitions from mono- to co- to tri-culture uniquely affect gene expression in breast cancer, stromal, and immune compartments. Biomed Microdevices. 2016;18: 70. pmid:27432323
  230. 230. Szot CS, Buchanan CF, Freeman JW, Rylander MN. In Vitro Angiogenesis Induced by Tumor-Endothelial Cell Co-Culture in Bilayered, Collagen I Hydrogel Bioengineered Tumors. Tissue Eng Part C Methods. 2013;19: 864–874. pmid:23516987
  231. 231. Gerhardt H, Golding M, Fruttiger M, Ruhrberg C, Lundkvist A, Abramsson A, et al. VEGF guides angiogenic sprouting utilizing endothelial tip cell filopodia. J Cell Biol. 2003;161: 1163–1177. pmid:12810700
  232. 232. Ye J, Wu D, Wu P, Chen Z, Huang J. The cancer stem cell niche: Cross talk between cancer stem cells and their microenvironment. Tumor Biol. 2014;35: 3945–3951. pmid:24420150
  233. 233. Lu J, Ye X, Fan F, Xia L, Bhattacharya R, Bellister S, et al. Endothelial Cells Promote the Colorectal Cancer Stem Cell Phenotype Through a Soluble Form of Jagged-1. Cancer Cell. 2014;23: 171–185.
  234. 234. Moraes C, Mehta G, Lesher-Perez SC, Takayama S. Organs-on-a-Chip: A Focus on Compartmentalized Microdevices. Ann Biomed Eng. 2012;40: 1211–1227. pmid:22065201
  235. 235. Mehta K, Mehta G, Takayama S, Linderman J. Quantitative inference of cellular parameters from microfluidic cell culture systems. Biotechnol Bioeng. 2009;103: 966–974. pmid:19388086
  236. 236. Mehta G, Mehta K, Sud D, Song JW, Bersano-Begey T, Futai N, et al. Quantitative measurement and control of oxygen levels in microfluidic poly(dimethylsiloxane) bioreactors during cell culture. Biomed Microdevices. 2007;9: 123–134. pmid:17160707
  237. 237. Kurtcuoglu V, Scholz CC, Marti HH, Hoogewijs D. Frequently asked questions in hypoxia research. 2015; 35–43.
  238. 238. McCord AM, Jamal M, Shankavarum UT, Lang FF, Camphausen K, Tofilon PJ. Physiologic Oxygen Concentration Enhances the Stem-Like Properties of CD133+ Human Glioblastoma Cells In vitro. Mol Cancer Res MCR. 2009;7: 489–497. pmid:19372578
  239. 239. Tong W-W, Tong G-H, Liu Y. Cancer stem cells and hypoxia-inducible factors. Int J Oncol. 2018;53: 469-. pmid:29845228
  240. 240. Wang J-Z, Zhu Y-X, Ma H-C, Chen S-N, Chao J-Y, Ruan W-D, et al. Developing multi-cellular tumor spheroid model (MCTS) in the chitosan/collagen/alginate (CCA) fibrous scaffold for anticancer drug screening. Mater Sci Eng C. 2016;62: 215–225. pmid:26952417
  241. 241. Feist P, Sun L, Liu X, Dovichi NJ, Hummon AB. Bottom-up proteomic analysis of single HCT 116 colon carcinoma multicellular spheroids. Rapid Commun Mass Spectrom RCM. 2015;29: 654–658. pmid:26212283
  242. 242. Dubiak-Szepietowska M, Karczmarczyk A, Jönsson-Niedziółka M, Winckler T, Feller K-H. Development of complex-shaped liver multicellular spheroids as a human-based model for nanoparticle toxicity assessment in vitro. Toxicol Appl Pharmacol. 2016;294: 78–85. pmid:26825373
  243. 243. Singh B, Cook KR, Vincent L, Hall CS, Martin C, Lucci A. Role of COX-2 in tumorospheres derived from a breast cancer cell line. J Surg Res. 2011;168: e39–e49. pmid:20462604
  244. 244. Zhao W, Luo Y, Li B, Zhang T. Tumorigenic lung tumorospheres exhibit stem-like features with significantly increased expression of CD133 and ABCG2. Mol Med Rep. 2016;14: 2598–2606. pmid:27432082
  245. 245. Klameth L, Rath B, Hochmaier M, Moser D, Redl M, Mungenast F, et al. Small cell lung cancer: model of circulating tumor cell tumorospheres in chemoresistance. Sci Rep. 2017;7. pmid:28706293
  246. 246. Cheng C-C, Chang J, Huang SC-C, Lin H-C, Ho A-S, Lim K-H, et al. YM155 as an inhibitor of cancer stemness simultaneously inhibits autophosphorylation of epidermal growth factor receptor and G9a-mediated stemness in lung cancer cells. PLoS ONE. 2017;12. pmid:28787001
  247. 247. Cheng C-C, Liao P-N, Ho A-S, Lim K-H, Chang J, Su Y-W, et al. STAT3 exacerbates survival of cancer stem-like tumorspheres in EGFR-positive colorectal cancers: RNAseq analysis and therapeutic screening. J Biomed Sci. 2018;25. pmid:30068339
  248. 248. Bahmad HF, Cheaito K, Chalhoub RM, Hadadeh O, Monzer A, Ballout F, et al. Sphere-Formation Assay: Three-Dimensional in vitro Culturing of Prostate Cancer Stem/Progenitor Sphere-Forming Cells. Front Oncol. 2018;8. pmid:30211124
  249. 249. Young SR, Saar M, Santos J, Nguyen HM, Vessella RL, Peehl DM. Establishment and serial passage of cell cultures derived from LuCaP xenografts. The Prostate. 2013;73: 1251–1262. pmid:23740600
  250. 250. Kondo J, Endo H, Okuyama H, Ishikawa O, Iishi H, Tsujii M, et al. Retaining cell-cell contact enables preparation and culture of spheroids composed of pure primary cancer cells from colorectal cancer. Proc Natl Acad Sci U S A. 2011;108: 6235–6240. pmid:21444794
  251. 251. Kwak J, Shim J-K, Kim DS, Lee J-H, Choi J, Park J, et al. Isolation and characterization of tumorspheres from a recurrent pineoblastoma patient: Feasibility of a patient-derived xenograft. Int J Oncol. 2016;49: 569–578. pmid:27277549
  252. 252. Jenkins RW, Aref AR, Lizotte PH, Ivanova E, Stinson S, Zhou CW, et al. Ex Vivo Profiling of PD-1 Blockade Using Organotypic Tumor Spheroids. Cancer Discov. 2018;8: 196–215. pmid:29101162
  253. 253. Eckhardt B, Gagliardi M, Iles LK, Evans K, Ivan C, Liu X, et al. Clinically relevant inflammatory breast cancer patient-derived xenograft–derived ex vivo model for evaluation of tumor-specific therapies. PLoS One. 2018;13: e0195932. pmid:29768500
  254. 254. Dolznig H, Rupp C, Puri C, Haslinger C, Schweifer N, Wieser E, et al. Modeling Colon Adenocarcinomas in Vitro: A 3D Co-Culture System Induces Cancer-Relevant Pathways upon Tumor Cell and Stromal Fibroblast Interaction. Am J Pathol. 2011;179: 487–501. pmid:21703426
  255. 255. Trapp V, Parmakhtiar B, Papazian V, Willmott L, Fruehauf JP. Anti-angiogenic effects of resveratrol mediated by decreased VEGF and increased TSP1 expression in melanoma-endothelial cell co-culture. Angiogenesis. 2010;13: 305–315. pmid:20927579
  256. 256. Amann A, Zwierzina M, Gamerith G, Bitsche M, Huber JM, Vogel GF, et al. Development of an Innovative 3D Cell Culture System to Study Tumour—Stroma Interactions in Non-Small Cell Lung Cancer Cells. PLOS ONE. 2014;9: e92511. pmid:24663399
  257. 257. Tsai S, McOlash L, Palen K, Johnson B, Duris C, Yang Q, et al. Development of primary human pancreatic cancer organoids, matched stromal and immune cells and 3D tumor microenvironment models. BMC Cancer. 2018;18. pmid:29587663
  258. 258. Bar-Ephraim YE, Kretzschmar K, Asra P, de Jongh E, Boonekamp KE, Drost J, et al. Modelling cancer immunomodulation using epithelial organoid cultures. bioRxiv. 2018;
  259. 259. Lindemans CA, Calafiore M, Mertelsmann AM, O’Connor MH, Dudakov JA, Jenq RR, et al. Interleukin-22 promotes intestinal-stem-cell-mediated epithelial regeneration. Nature. 2015;528: 560–564. pmid:26649819
  260. 260. Zhang H, Hao C, Wang Y, Ji S, Zhang X, Zhang W, et al. Sohlh2 inhibits human ovarian cancer cell invasion and metastasis by transcriptional inactivation of MMP9. Mol Carcinog. 2016;55: 1127–1137. pmid:26153894
  261. 261. Marsano A, Conficconi C, Lemme M, Occhetta P, Gaudiello E, Votta E, et al. Beating heart on a chip: a novel microfluidic platform to generate functional 3D cardiac microtissues. Lab Chip. 2016;16: 599–610. pmid:26758922
  262. 262. Bavli D, Prill S, Ezra E, Levy G, Cohen M, Vinken M, et al. Real-time monitoring of metabolic function in liver-on-chip microdevices tracks the dynamics of mitochondrial dysfunction. Proc Natl Acad Sci. 2016;113: E2231–E2240. pmid:27044092
  263. 263. Hao S, Ha L, Cheng G, Wan Y, Xia Y, Sosnoski DM, et al. A Spontaneous 3D Bone-On-a-Chip for Bone Metastasis Study of Breast Cancer Cells. Small. 2018;14: 1702787. pmid:29399951
  264. 264. Huh D (Dan). A Human Breathing Lung-on-a-Chip. Ann Am Thorac Soc. 2015;12: S42–S44. pmid:25830834
  265. 265. Kim HJ, Huh D, Hamilton G, Ingber DE. Human gut-on-a-chip inhabited by microbial flora that experiences intestinal peristalsis-like motions and flow. Lab Chip. 2012;12: 2165–2174. pmid:22434367
  266. 266. Sobrino A, Phan DTT, Datta R, Wang X, Hachey SJ, Romero-López M, et al. 3D microtumors in vitro supported by perfused vascular networks. Sci Rep. 2016;6. pmid:27549930
  267. 267. Hsiao AY, Torisawa Y, Tung Y-C, Sud S, Taichman RS, Pienta KJ, et al. Microfluidic system for formation of PC-3 prostate cancer co-culture spheroids. Biomaterials. 2009;30: 3020–3027. pmid:19304321
  268. 268. Zhang B, Korolj A, Lai BFL, Radisic M. Advances in organ-on-a-chip engineering. Nat Rev Mater. 2018;3: 257.
  269. 269. Beckwitt CH, Clark AM, Wheeler S, Taylor DL, Stolz DB, Griffith L, et al. Liver ‘organ on a chip’. Exp Cell Res. 2018;363: 15–25. pmid:29291400
  270. 270. Wilmer MJ, Ng CP, Lanz HL, Vulto P, Suter-Dick L, Masereeuw R. Kidney-on-a-Chip Technology for Drug-Induced Nephrotoxicity Screening. Trends Biotechnol. 2016;34: 156–170. pmid:26708346
  271. 271. Skardal A, Devarasetty M, Forsythe S, Atala A, Soker S. A reductionist metastasis-on-a-chip platform for in vitro tumor progression modeling and drug screening. Biotechnol Bioeng. 2016;113: 2020–2032. pmid:26888480
  272. 272. Tsai H-F, Trubelja A, Shen AQ, Bao G. Tumour-on-a-chip: microfluidic models of tumour morphology, growth and microenvironment. J R Soc Interface. 2017;14. pmid:28637915
  273. 273. Ahn J, Sei YJ, Jeon NL, Kim Y. Tumor Microenvironment on a Chip: The Progress and Future Perspective. Bioengineering. 2017;4. pmid:28952543
  274. 274. Hachey SJ, Hughes CCW. Applications of tumor chip technology. Lab Chip. 2018;18: 2893–2912. pmid:30156248
  275. 275. Crespo I, Coukos G, Doucey M-A, Xenarios I. Modelling approaches in tumor microenvironment. J Cancer Immunol Ther. 2018;1. http://www.alliedacademies.org/abstract/modelling-approaches--in--tumor-microenvironment-9508.html
  276. 276. Altrock PM, Liu LL, Michor F. The mathematics of cancer: integrating quantitative models. Nat Rev Cancer. 2015;15: 730–745. pmid:26597528
  277. 277. Mahlbacher G, Curtis LT, Lowengrub J, Frieboes HB. Mathematical modeling of tumor-associated macrophage interactions with the cancer microenvironment. J Immunother Cancer. 2018;6. pmid:29382395
  278. 278. Mumenthaler SM, Foo J, Choi NC, Heise N, Leder K, Agus DB, et al. The Impact of Microenvironmental Heterogeneity on the Evolution of Drug Resistance in Cancer Cells. Cancer Inform. 2015;14: 19–31. pmid:26244007
  279. 279. Anderson ARA. A hybrid mathematical model of solid tumour invasion: the importance of cell adhesion. Math Med Biol J IMA. 2005;22: 163–186. pmid:15781426
  280. 280. d’Onofrio A. Metamodeling tumor–immune system interaction, tumor evasion and immunotherapy. Math Comput Model. 2008;47: 614–637.
  281. 281. Wilkie K, Hahnfeldt P, Hlatky L. Using Ordinary Differential Equations to Explore Cancer-Immune Dynamics and Tumor Dormancy. bioRxiv. 2016; 049874.
  282. 282. Webb SD, Sherratt JA, Fish RG. Cells behaving badly: a theoretical model for the Fas/FasL system in tumour immunology. Math Biosci. 2002;179: 113–129. pmid:12208611
  283. 283. Tomasetti C, Vogelstein B, Parmigiani G. Half or more of the somatic mutations in cancers of self-renewing tissues originate prior to tumor initiation. Proc Natl Acad Sci U S A. 2013;110: 1999–2004. pmid:23345422
  284. 284. Lee H-O, Silva AS, Concilio S, Li Y-S, Slifker M, Gatenby RA, et al. Evolution of Tumor Invasiveness: The Adaptive Tumor Microenvironment Landscape Model. Cancer Res. 2011;71: 6327–6337. pmid:21859828
  285. 285. Haeno H, Gonen M, Davis MB, Herman JM, Iacobuzio-Donahue CA, Michor F. Computational modeling of pancreatic cancer reveals kinetics of metastasis suggesting optimum treatment strategies. Cell. 2012;148: 362–375. pmid:22265421
  286. 286. Xu J, Vilanova G, Gomez H. A Mathematical Model Coupling Tumor Growth and Angiogenesis. PLOS ONE. 2016;11: e0149422. pmid:26891163
  287. 287. Stefanini MO, Qutub AA, Gabhann FM, Popel AS. Computational models of VEGF-associated angiogenic processes in cancer. Math Med Biol J IMA. 2012;29: 85–94. pmid:21266494
  288. 288. Harpold HLP, Alvord EC, Swanson KR. The evolution of mathematical modeling of glioma proliferation and invasion. J Neuropathol Exp Neurol. 2007;66: 1–9. pmid:17204931
  289. 289. Böttcher MA, Held-Feindt J, Synowitz M, Lucius R, Traulsen A, Hattermann K. Modeling treatment-dependent glioma growth including a dormant tumor cell subpopulation. BMC Cancer. 2018;18: 376. pmid:29614985
  290. 290. Goldie JH, Coldman AJ. A mathematic model for relating the drug sensitivity of tumors to their spontaneous mutation rate. Cancer Treat Rep. 1979;63: 1727–1733. pmid:526911
  291. 291. Bray LJ, Binner M, Körner Y, von Bonin M, Bornhäuser M, Werner C. A three-dimensional ex vivo tri-culture model mimics cell-cell interactions between acute myeloid leukemia and the vascular niche. Haematologica. 2017;102: 1215. pmid:28360147
  292. 292. Lamble AJ, Lind EF. Targeting the Immune Microenvironment in Acute Myeloid Leukemia: A Focus on T Cell Immunity. Front Oncol. 2018;8. pmid:29951373
  293. 293. Pease JC, Brewer M, Tirnauer JS. Spontaneous spheroid budding from monolayers: a potential contribution to ovarian cancer dissemination. Biol Open. 2012;1: 622–628. pmid:23213456
  294. 294. Meng E, Mitra A, Tripathi K, Finan MA, Scalici J, McClellan S, et al. ALDH1A1 Maintains Ovarian Cancer Stem Cell-Like Properties by Altered Regulation of Cell Cycle Checkpoint and DNA Repair Network Signaling. PLoS ONE. 2014;9. pmid:25216266
  295. 295. Ishiguro T, Sato A, Ohata H, Ikarashi Y, Takahashi R, Ochiya T, et al. Establishment and Characterization of an In Vitro Model of Ovarian Cancer Stem-like Cells with an Enhanced Proliferative Capacity. Cancer Res. 2016;76: 150–160. pmid:26669863
  296. 296. Kim G-A, Ginga NJ, Takayama S. Integration of Sensors in Gastrointestinal Organoid Culture for Biological Analysis. Cell Mol Gastroenterol Hepatol. 2018;6: 123–131.e1. pmid:29928682
  297. 297. Personalized Proteome Profiles of Healthy and Tumor Human Colon Organoids Reveal Both Individual Diversity and Basic Features of Colorectal Cancer. Cell Rep. 2017;18: 263–274. pmid:28052255
  298. 298. Roda A, Michelini E, Caliceti C, Guardigli M, Mirasoli M, Simoni P. Advanced bioanalytics for precision medicine. Anal Bioanal Chem. 2018;410: 669–677. pmid:29026940