Use of atomic force microscopy to assess the biomechanical properties of 3D tumor cell models

Cover Page


Cite item

Full Text

Open Access Open Access
Restricted Access Access granted
Restricted Access Subscription or Fee Access

Abstract

Multicellular spheroids are a unique object model for toxicological studies. Cells in a three-dimensional (3D) cluster contain a microenvironment and intercellular communication, which allows spheroids to be used as a more realistic model than traditional cell cultures. Tumor microregions consist of heterogeneous populations of cancer cells, in which cell growth and response to antitumor drugs depend on their 3D architecture, intercellular contacts, and interaction with the microenvironment. Tumor growth and progression are also strongly influenced by mechanical cues. Currently, 3D cell culture models are powerful tools for studying the toxicity of drug compounds and nanomaterials of different compositions and morphologies.

This review presents data on the use of various techniques, particularly atomic force microscopy, to investigate changes in the mechanical properties of cells in spheroids. Specifically, the use of atomic force microscopy as a tool to reveal physicochemical parameters of cells during pathophysiological processes or drug exposure is considered. The relevance of this review is attributed to the increasing interest in the role of biomechanical properties of tissues, cells, and subcellular structures as markers of pathophysiological conditions.

Full Text

INTRODUCTION

A tumor is believed as not just a collection of cells capable of high proliferation but a separate tissue that is distinguished by features of development and structure of the circulatory system, particularly associated immune cells and fibroblasts. Tumor resistance to therapy was linked to the properties of the extracellular matrix (ECM) [1]. Tumor cells within a solid tumor have different physiological states because of different accesses to oxygen and nutrients [2]. A comprehensive assessment of the morphology and physicochemical parameters of the tumor parenchyma and stroma has application and fundamental importance [3]. Tumor microregions consist of heterogeneous populations of cancer cells, in which cell growth and response to antitumor drugs depend on their 3D architecture, intercellular contacts, and interaction with the microenvironment. Tumor growth and progression are also strongly influenced by mechanical cues [4]. Thus, the state of cells within the tumor tissue must be examined to predict the disease course and assess the efficacy of the applied antitumor therapy. Owing to the complexity of studying the mechanisms of tumor development and their response to drug therapy in vivo, several solid tumor models have been developed, which reproduce the properties of tumor tissue to varying degrees [5]. One such model is 3D spheroids, which are multicellular 3D aggregates of tumor cells that reproduce the cell–cell and cell–intracellular matrix interactions found in solid tumors. 3D spheroids represent the most adequate model for examining the physical and mechanical properties of cells and preclinical studies of therapeutic drugs [6]; thus, they are increasingly used as models for basic research and drug screening [7]. Multicellular tumor spheroids (MCOs) are emerging as a powerful in vitro model for preclinical experiments, closely resembling the three-dimensional (3D) structure of avascularized tumors in vivo. Multicellular spheroids are an important in vitro model for preclinical experiments because they mimic the 3D structure of avascularized tumors in vivo [8]. When the diameter of multicellular clusters is increased to >500 μm, they can reproduce various pathophysiological features of living human solid tumors [9] because they exhibit altered gene and protein expression profiles, representing an intermediate model between two-dimensional (2D) and in vivo models [10]. However, differences in drug sensitivity have been observed between cells in monolayer and multicellular spheroids [11]. Moreover, methodological approaches to characterize the complex 3D multicellular organization of spheroids are being actively developed [12]. This review will closely investigate the mechanical properties of cells within 3D spheroids, with particular emphasis on the use of atomic force microscopy (AFM).

For example, it is reported that 3D spheroids demonstrated stronger anticancer drug resistance than 2D analogs [3]. Tumor cells inside a solid tumor differ in their physiological state, which is associated with different accesses to oxygen and nutrients [4]. A comprehensive assessment of the morphology and physicochemical parameters of the tumor parenchyma and stroma has a special application and fundamental importance [5]. Tumor microregions consist of heterogeneous populations of cancer cells in which cell growth and response to antitumor drugs depend on their 3D architecture, intercellular contacts, and interaction with the microenvironment. Tumor growth and progression are also strongly influenced by mechanical signals [6].

The state of cells inside the tumor tissue must be examined to predict the disease course and evaluate the effectiveness of the antitumor therapy. Owing to the complexity of studying the mechanisms of tumor development and their response to drug therapy in vivo, several solid tumor models have been developed, which reproduce the properties of tumor tissue to varying degrees [7]. One such model is 3D spheroids, which are multicellular 3D aggregates of tumor cells that reproduce the cell–cell and cell–ECM interactions in solid tumors [8]. 3D spheroids represent the most adequate model for studying the physicomechanical properties of cells and preclinical studies of therapeutic drugs [8], and the use of 3D spheroids as models for basic research and screening of medicines has been growing recently [9–11]. MCOs are becoming powerful in vitro models for preclinical experiments [12]. MCOs have intercellular contacts and cell–ECM interactions, which are necessary for the regulation of cell behavior and function. If the diameter of the MCOs increased to >500 microns, MCOs can reproduce many pathophysiological features of living solid human tumors [13], and in comparison with conventional 2D cell culture models, they demonstrated altered gene and protein expression profiles, which allows them to be considered an intermediate model between a 2D monolayer and solid tumors in vivo. Given these properties, MCOs are currently used in a wide range of cell biology research, bridging the gap between preclinical and clinical results [14]. Moreover, methodological approaches are actively developing to characterize the complex 3D multicellular organization of spheroids [15].

The review provides the latest data on the changes in the biomechanical properties of cells in the composition of 3D structures under the action of medicinal compounds. The mechanical properties of cells in the composition of 3D spheroids were also examined using various methods, with special emphasis on the use of AFM.

DIFFERENCE BETWEEN THE BIOMECHANICAL PROPERTIES OF CANCER CELLS AND NONCANCER CELLS IN 2D AND 3D MODELS

Even at the initial stages of carcinogenesis, changes in the mechanical and morphological properties of cells occur, which can be used as disease markers [16]. Currently, the role of biomechanical properties of tissues, cells, and subcellular structures is receiving research interest as markers of pathophysiological conditions [17, 18]. As known, different cells have different stiffness, for example, the Young modulus of embryonic stem cells is 500 Pa; osteoblasts, 1.6–2.6 kPa; and osteosarcoma cells, 1 kPa [19, 20]. Nontumor cells tend to be tougher than cancer cells [21, 22]. The reason for the observed biomechanical differences is not yet fully clear; however, during the transformation of healthy cells into cancerous ones, the structure of their cytoskeleton changes from an organized to an irregular network, which probably changes their mechanical properties [23]. Biomechanical properties such as cell adhesion and cytosolic viscosity also differ in tumor and normal cells [21]. Moreover, the tumor is generally tougher than the surrounding healthy tissues [17], and the rigidity of cancer cells increases with the induction of apoptosis [24].

Traditionally, cells growing as a 2D monolayer have been used to examine the effectiveness of antitumor drugs (including AFM). However, in the last decade, the development of methods for growing 3D cell cultures has significantly changed the way preclinical drug research is conducted. Monolayer cell cultures are believed to not reproduce the mechanisms of drug resistance of tumors associated with the tumor cell microenvironment [25–27], and the smallest functional units of tissues preserving the basic physiological characteristics are micro-sized tissues in the range of 100 microns to 1 mm [28]. In 2D cell cultures, cell–cell and cell–ECM interactions are reduced, and the cellular reactivity level is limited [29]. Moreover, the cell culture medium can influence the cell phenotype and, consequently, affect the cell response to added substances, such as drugs [30]. The phenotype and functions of each cell largely depend on carefully thought-out interactions with neighboring cells, ECM, and proteins [31]. These cell–cell and cell–ECM interactions differ in 2D and 3D cultures and between cell layers in spheroidal structures, and this may affect cytotoxicity [32]. Thus, testing the toxicity of materials and substances on 2D cell cultures does not allow for the accurate prediction of body changes. 3D cell cultures more accurately mimic the natural cellular microenvironment. The morphology and physiology of cells in 3D cultures differ from that of cells in 2D cultures, demonstrating reactions that in some ways more resemble in vivo behavior [33, 34]. For example, in 2D models of Alzheimer’s disease, medium removal will mean that the secreted beta-amyloid (Aß) will be removed; therefore, the analysis of Aß aggregation will change, and the use of 3D cell culture may limit the diffusion of Aß into the culture medium [29]. In addition, the use of 3D spheroids as tumor models allows us to approach in vivo conditions because, as mentioned above, they retain the influence of the ECM, intercellular interactions, and tissue-like 3D organization [33].

Various methods of 3D modeling of tumors (cell culture on special hydrogels and substrates, in micro- and macro-bioreactors, production of spheroids and organoids, organs-on-a-chip, and computer and bioinformatic models) are available, which reproduce the properties of solid tumors to varying degrees [34, 35]. 3D cell cultures in the form of 3D cell spheroids are one of the most reliable and widely used tumor models because they reproduce numerous tumor properties, including oxygen gradient, intercellular interactions, and cell–ECM interactions observed in solid tumors [8, 12, 34, 36, 37]. In addition, spheroids can grow up to several hundred micrometers in diameter and exhibit a cell proliferation gradient similar to that observed in tumor microblasts [38]. Like tumors, the cellular organization inside the spheroid varies and depends on the stage of spheroid growth [39, 40]. Loosely located cells on the surface of the spheroid demonstrated the highest level of proliferation, and cells located in internal compact areas are in a state of hypoxia or anoxia (spheroids with a diameter of 250–300 microns) [41] or necrosis (spheroids with a diameter of >550 microns) [39]. Given the importance of spheroids as in vitro tumor models in preclinical studies, their physical and mechanical properties must be elucidated using several methodological approaches.

COMPARATIVE ANALYSIS OF DIFFERENT APPROACHES FOR THE DETECTION OF BIOMECHANICAL PROPERTIES OF CELLS

Optical coherent elastography has been used to measure the biomechanical properties of spheroids from breast cancer cell lines with and without multidrug resistance [42]. Spheroids from multidrug-resistant cells are tougher than spheroids from drug-sensitive cells. In addition, the stiffness of drug-sensitive spheroids increased after their 24-h co-cultivation with extracellular vesicles isolated from multidrug-resistant cells [42, 43]. 3D light microscopy in combination with mathematical modeling has recently been proposed to estimate the elasticity of multicellular spheroids [44]. Previously, Kelvin probe force microscopy was also used for the quantitative analysis of biomolecular interactions [45] by measuring the contact potential difference between the probe and the sample; however, the resolution of the AFM is much higher [46].

Recently, AFM has been considered a tool for diagnosing diseases and evaluating the effectiveness of treatment based on differences in the topography and mechanical properties of cells. Some serious human diseases indicate changes in the ultrastructure of blood cell membranes. Chen found that mononuclear cells of the peripheral blood of patients with uremia had a larger volume, height, and roughness of the membrane than cells of healthy people [47]. AFM has also been used to detect morphological changes in erythrocytes in type 2 diabetes [48], multiple myeloma [49], and iron deficiency anemia [50]. The AFM also allowed for real-time examination of the dynamics of structural changes associated with neurodegeneration induced by the activation of N-methyl-D-aspartate receptors (a subgroup of glutamate receptors) in living SH-SY5Y cells [51].

The reaction of cancer cells to drug therapy is also being actively explored. Several studies have shown an increase in the roughness of cancer cells after the induction of apoptosis using drugs [52–56]. A time- and dose-dependent increase in the roughness of the membranes of HeLa, HepG2, and C6 cells after treatment with colchicine and citrabine was reported [57]. The rigidity of cells and tissues (which can be judged by Young’s modulus) has been identified as a key factor in cell function, including their adhesion, mobility, and invasion abilities [58]. Cancer cells are on average softer (exhibit a lower Young’s modulus) than noncancerous cells; thus, measuring stiffness can help distinguish between malignant and nonmalignant tissue [59] and evaluate the effectiveness of drug therapy. In turn, literature analysis allows us to consider AFM as a promising method for such studies, including the study of multicellular clusters.

ADVANTAGES OF USING ATOMIC FORCE MICROSCOPY IN ASSESSING CHANGES IN THE BIOMECHANICAL PROPERTIES OF CELLS IN 3D SPHEROIDS

The AFM, originally developed for imaging various surfaces, is currently actively used to examine the mechanical properties of biological objects, including tissues, cells, membranes, protein complexes, individual protein molecules and nucleic acids [60], as in sections of fixed tissues [61], and physiological conditions [62, 63]. The use of AFM for the integral assessment of the mechanical properties of an entire 3D spheroid has been described [64, 65]; however, the physiological state of cells on the surface and inside the spheroids and mechanical properties of cells on varying distances from the surface of the spheroid also differ. AFM is considered a fast and convenient method for detecting such changes. However, when studying cells in a monolayer, the use of various substrates (matrix proteins, gels, etc.) can strongly affect the measured parameters, and nonphysiological conditions cast doubt on the reliability of the results obtained. To assess the rheological properties of spheroids, methods such as centrifugation, compression with parallel plates, aspiration with a micropipette, and rupture of connected spheroids were also used [64, 66, 67]. Although these approaches provide some information about the mechanical properties of large biological systems, they have several limitations. For example, the pipette aspiration method requires optical correction of the focusing error of the micropipette and inner diameter and measurement and compensation of the taper of the micropipette [68], whereas microplate rheology does not have sufficient force sensitivity for a thorough description of biomechanical phenomena [69]. In contrast to these approaches, AFM allows for the direct application and measurement of forces in the range from pico- to micronewton in spatially strictly defined areas with precise vertical positioning control in the range from subnanometers to several tens of micrometers.

The AFM system is based on a flexible cantilever, and its movement is tracked by the reflection of a laser beam directed at the outer surface of the cantilever using a photodetector. The resolution of AFM images can reach 0.5–1.0 nm in the plane and 0.1–0.2 nm in height, and the recorded forces range from piconewtons to micronewtons. The fixation and dehydration of cell samples for AFM are significantly less destructive than sample preparation of samples for transmission and scanning electron microscopies [70]. Using AFM, the choice of cantilever geometry plays a special role when studying the mechanical properties of cells. For example, cantilevers with a sharp tip can be used to probe nanoscale regions of the cell surface, whereas flat cantilevers without a tip are more suitable for examining the mechanical properties of an entire cell [71]. In the AFM analysis of the mechanical properties of cells [72, 73] and intracellular structures [74], pressure or compression of the cell with standard or modified AFM probes is used. Such an application of AFM is similar to the method of compressing objects with parallel microplates [75], where a cantilever without a tip is used as an upper microplate for performing cell deformation and measuring forces [76, 77]. An obstacle to the widespread use of AFM in cell research is the standard cantilever tilt of 8–12, which does not allow for the application of a unidirectional load to the sample surface. To solve this problem, a modification of the cantilever without a tip was proposed to create an adhesive wedge-shaped plate on it, allowing for the application of a uniform load to the sample [78].

In the study of spheroids by AFM, in spheroids, as in solid tumors, gradual increases in pressure are induced by the intrinsic growth of the spheroid/tumor [68]. Growth-induced pressure affects tumor growth and its resistance to therapy. In addition, the physical properties of spheroids are influenced by the conditions of their formation: even if spheroids have the same size, the growth-induced pressure in spheroids grown for 2 days from 5,000 cells was lower than the pressure in spheroids grown for 6 days from 500 cells [68]. Young’s modulus obtained in this study for spheroids from HCT116 cells was comparable to those previously found for spheroids from breast cancer cells and healthy cells using micro-tweezers [79]. In large cells or 3D multicellular aggregates, standard AFM probes may be unsuitable because their small size limits them to press only on a single cell or part of it. Although the size of AFM probes can be increased by gluing large beads with a diameter of up to 50 microns to the cantilever, this approach does not allow for achieving a uniform and stable contact area for homogeneous and reproducible compression of a large biological system (>100 microns). In addition, the shape of the pressure device begins to play an important role in such measurements [64]. In this regard, a method has been proposed for creating flat AFM macro-probes consisting of a large flat plate connected to a chip using two flexible legs, which are used to apply/measure compression forces similar to standard AFM cantilevers. With the help of such macrobonds, the viscoelastic properties of small (up to 200 microns in diameter) and large (>200 microns in diameter) spheroids from T47D tumor cells were evaluated, and in large spheroids, the outer layer of cells was found to be mainly subjected to compression deformation, whereas small spheroids are exposed to much deeper deformations, which corresponds to data on the similar cellular organization of small spheroids and the outer layer of large spheroids [64].

CONCLUSION

Multicellular clusters (spheroids) are important in vitro models for preclinical experiments, and with increased diameter, they can reproduce many pathophysiological features of human solid tumors. 3D spheroids represent an intermediate model between 2D and in vivo models. Differences in drug sensitivity between monolayer cells and multicellular spheroids are observed, with important changes occurring at the level of nanomechanical properties of cells, such as roughness, stiffness, etc. In addition, methodological approaches for characterizing the complex 3D multicellular organization of spheroids are being actively developed. In this literature review, current data on methods for detecting changes in mechanical properties of cells in 3D spheroids are considered, and atomic force microscopy application is thoroughly discussed.

ДОПОЛНИТЕЛЬНАЯ ИНФОРМАЦИЯ

Источник финансирования. Авторы заявляют об отсутствии внешнего финансирования при проведении исследования.

Конфликт интересов. Авторы декларируют отсутствие явных и потенциальных конфликтов интересов, связанных с публикацией настоящей статьи.

Вклад авторов. Все авторы подтверждают соответствие своего авторства международным критериям ICMJE (все авторы внесли существенный вклад в разработку концепции, проведение исследования и подготовку статьи, прочли и одобрили финальную версию перед публикацией). Наибольший вклад распределён следующим образом: К.С. Пучкова — обзор литературы, написание текста и редактирование статьи; В.Р. Лопарева — обзор литературы, сбор и анализ литературных источников, написание введения статьи; Е.В. Шепелева — сбор и анализ литературных источников, подготовка и написание заключения; О.А. Магомедова — сбор литературных источников, оформление литературных ссылок; Д.В. Иванова — анализ литературных источников; Ю.В. Замская — написание текста и редактирование статьи.

ADDITIONAL INFORMATION

Funding source. This study was not supported by any external sources of funding.

Competing interests. The authors declare that they have no competing interests.

Authors’ contribution. All authors confirm that their authorship meets the international ICMJE criteria (all authors have made a significant contribution to the development of the concept, research and preparation of the article, read and approved the final version before publication). The greatest contribution is distributed as follows: K.S. Puchkova — literature review, writing the text and editing the article; V.R. Lopareva — literature review, collection and analysis of literary sources, writing the introduction of the article; E.V. Shepeleva — collection and analysis of literary sources, preparation and writing the conclusion; O.A. Magomedova — collection of literary sources, design of literary references; D.V. Ivanova — analysis of literary sources; Y.V. Zamskaya — writing the text and editing the article.

×

About the authors

Ksenia S. Puchkova

N.I. Pirogov Russian National Research Medical University

Email: ms.puchkova@yandex.ru
ORCID iD: 0009-0004-1820-3209
Russian Federation, Moscow

Valeria R. Lopareva

N.I. Pirogov Russian National Research Medical University

Email: gontar_78@mail.ru
ORCID iD: 0009-0007-9195-5310
Russian Federation, Moscow

Ekaterina V. Shepeleva

N.I. Pirogov Russian National Research Medical University

Email: plokatyayay@mail.ru
ORCID iD: 0009-0005-9628-7492
Russian Federation, Moscow

Oksana A. Magomedova

N.I. Pirogov Russian National Research Medical University

Email: oksana.mag1998@mail.ru
ORCID iD: 0009-0006-8292-7951
Russian Federation, Moscow

Daria V. Ivanova

N.I. Pirogov Russian National Research Medical University

Email: dar.ivanova22@gmail.com
ORCID iD: 0009-0001-0616-2444
Russian Federation, Moscow

Yulia V. Zamskaya

N.I. Pirogov Russian National Research Medical University

Author for correspondence.
Email: Y.Fom00@mail.ru
ORCID iD: 0009-0005-8689-4029
Russian Federation, Moscow

References

  1. Gkretsi V, Stylianou A, Papageorgis P, et al. Remodeling components of the tumor microenvironment to enhance cancer therapy. Front Oncol. 2015;5:214. doi: 10.3389/fonc.2015.00214
  2. Walker C, Mojares E, Del Río Hernández A. Role of extracellular matrix in development and cancer progression. Int J Mol Sci. 2018;19(10):3028. doi: 10.3390/ijms19103028
  3. Perut F, Sbrana FV, Avnet S, et al. Spheroid-based 3D cell cultures identify salinomycin as a promising drug for the treatment of chondrosarcoma. J Orthop Res. 2018;36:2305–2312. doi: 10.1002/jor.23880
  4. Edmondson R, Adcock AF, Yang L. Influence of matrices on 3D-cultured prostate cancer cells’ drug response and expression of drug-action associated proteins. PLoS One. 2016;11(6):e0158116. doi: 10.1371/journal.pone.0158116
  5. Tarin D. Role of the host stroma in cancer and its therapeutic significance. Cancer Metastasis Rev. 2013;32(3-4):553–566. doi: 10.1007/s10555-013-9438-4
  6. Jain RK, Martin JD, Stylianopoulos T. The role of mechanical forces in tumor growth and therapy. Annu Rev Biomed Eng. 2014;16:321–346. doi: 10.1146/annurev-bioeng-071813-105259
  7. Bregenzer ME, Horst EN, Mehta P, et al. Tumor modeling maintains diverse pathology in vitro. Ann Transl Med. 2019;7(Suppl. 8). doi: 10.21037/atm.2019.12.32
  8. Hirschhaeuser F, Menne H, Dittfeld C, et al. Multicellular tumor spheroids: an underestimated tool is catching up again. J Biotechnol. 2010;148(1):3–15. doi: 10.1016/j.jbiotec.2010.01.012
  9. Sant S, Johnston PA. The production of 3D tumor spheroids for cancer drug discovery. Drug Discov Today Technol. 2017;23:27–36. doi: 10.1016/j.ddtec.2017.03.002
  10. Białkowska K, Komorowski P, Bryszewska M, Miłowska K. Spheroids as a type of three-dimensional cell cultures-examples of methods of preparation and the most important application. Int J Mol Sci. 2020;21(17):6225. doi: 10.3390/ijms21176225
  11. Pinto B, Henriques AC, Silva PMA, Bousbaa H. Three-dimensional spheroids as in vitro preclinical models for cancer research. Pharmaceutics. 2020;12(12):1186. doi: 10.3390/pharmaceutics12121186
  12. Zanoni M, Cortesi M, Zamagni A, et al. Modeling neoplastic disease with spheroids and organoids. J Hematol Oncol. 2020;13(1):97. doi: 10.1186/s13045-020-00931-0
  13. Han SJ, Kwon S, Kim KS. Challenges of applying multicellular tumor spheroids in preclinical phase. Cancer Cell Int. 2021;21(1):152. doi: 10.1186/s12935-021-01853-8
  14. Wen S, Tu X, Zang Q, et al. Liquid chromatography-mass spectrometry-based metabolomics and fluxomics reveals the metabolic alterations in glioma U87MG multicellular tumor spheroids versus two-dimensional cell cultures. Rapid Commun Mass Spectrom. 2024;38(2):e9670. doi: 10.1002/rcm.9670
  15. Desmaison A, Guillaume L, Triclin S, et al. Impact of physical confinement on nuclei geometry and cell division dynamics in 3D spheroids. Sci Rep. 2018;8(1):8785. doi: 10.1038/s41598-018-27060-6
  16. Lekka M, Gil D, Pogoda K, et al. Cancer cell detection in tissue sections using AFM. Arch Biochem Biophys. 2012;518(2):151–156. doi: 10.1016/j.abb.2011.12.013
  17. Stylianou A, Kontomaris SV, Grant C, Alexandratou E. Atomic force microscopy on biological materials related to pathological conditions. Scanning. 2019;2019:8452851. doi: 10.1155/2019/8452851
  18. Runel G, Lopez-Ramirez N, Chlasta J, Masse I. Biomechanical properties of cancer cells. Cells. 2021;10(4):887. doi: 10.3390/cells10040887
  19. Chowdhury F, Na S, Li D, et al. Material properties of the cell dictate stress-induced spreading and differentiation in embryonic stem cells. Nat Mater. 2010;9(1):82–88. doi: 10.1038/nmat2563
  20. Docheva D, Padula D, Popov C, et al. Researching into the cellular shape, volume and elasticity of mesenchymal stem cells, osteoblasts and osteosarcoma cells by atomic force microscopy. J Cell Mol Med. 2008;12(2):537–552. doi: 10.1111/j.1582-4934.2007.00138.x
  21. Suresh S. Biomechanics and biophysics of cancer cells. Acta Biomater. 2007;3(4):413–438. doi: 10.1016/j.actbio.2007.04.002
  22. Li QS, Lee GY, Ong CN, Lim CT. AFM indentation study of breast cancer cells. Biochem Biophys Res Commun. 2008;374(4):609–613. doi: 10.1016/j.bbrc.2008.07.078
  23. Nia HT, Munn LL, Jain RK. Physical traits of cancer. Science. 2020;370(6516):eaaz0868. doi: 10.1126/science.aaz0868
  24. Pi J, Cai J. Cell Topography and its quantitative imaging by AFM. Methods Mol Biol. 2019;1886:99–113. doi: 10.1007/978-1-4939-8894-5_6
  25. Vargo-Gogola T, Rosen JM. Modelling breast cancer: one size does not fit all. Nat Rev Cancer. 2007;7(9):659–672. doi: 10.1038/nrc2193
  26. Correia AL, Bissell MJ. The tumor microenvironment is a dominant force in multidrug resistance. Drug Resist Updat. 2012;15(1-2):39–49. doi: 10.1016/j.drup.2012.01.006
  27. Meads MB, Gatenby RA, Dalton WS. Environment-mediated drug resistance: a major contributor to minimal residual disease. Nat Rev Cancer. 2009;9(9):665–674. doi: 10.1038/nrc2714
  28. Bhatia SN, Chen CS. Tissue engineering at the micro-scale. Biomedical Microdevices. 1999;2:131–144. doi: 10.1023/A:1009949704750
  29. Centeno EGZ, Cimarosti H, Bithell A. 2D versus 3D human induced pluripotent stem cell-derived cultures for neurodegenerative disease modelling. Mol Neurodegener. 2018;13(1):27. doi: 10.1186/s13024-018-0258-4
  30. Goodman TT, Ng CP, Pun SH. 3-D tissue culture systems for the evaluation and optimization of nanoparticle-based drug carriers. Bioconjug Chem. 2008;19(10):1951–1959. doi: 10.1021/bc800233a
  31. Lee J, Lilly GD, Doty RC, et al. In vitro toxicity testing of nanoparticles in 3D cell culture. Small. 2009;5(10):1213–1221. doi: 10.1002/smll.200801788
  32. Moshksayan K, Kashaninejad N, Warkiani ME, et al. Spheroids-on-a-chip: Recent advances and design considerations in microfluidic platforms for spheroid formation and culture. Sensors and Actuators B: Chemical. 2018;263:151–176. doi: 10.1016/j.snb.2018.01.223
  33. 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(4):207–218. doi: 10.1089/adt.2014.573
  34. Bregenzer ME, Horst EN, Mehta P, et al. Integrated cancer tissue engineering models for precision medicine. PLoS One. 2019;14(5):e0216564. doi: 10.1371/journal.pone.0216564
  35. Kosheleva NV, Efremov YM, Koteneva PI, et al. Building a tissue: Mesenchymal and epithelial cell spheroids mechanical properties at micro- and nanoscale. Acta Biomater. 2023;165:140–152. doi: 10.1016/j.actbio.2022.09.051
  36. Nunes AS, Barros AS, Costa EC, et al. 3D tumor spheroids as in vitro models to mimic in vivo human solid tumors resistance to therapeutic drugs. Biotechnol Bioeng. 2019;116(1):206–226. doi: 10.1002/bit.26845
  37. Nath S, Devi GR. Three-dimensional culture systems in cancer research: Focus on tumor spheroid model. Pharmacol Ther. 2016;163:94–108. doi: 10.1016/j.pharmthera.2016.03.013
  38. Laurent J, Frongia C, Cazales M, et al. Multicellular tumor spheroid models to explore cell cycle checkpoints in 3D. BMC Cancer. 2013;13:73. doi: 10.1186/1471-2407-13-73
  39. Riffle S, Pandey RN, Albert M, Hegde RS. Linking hypoxia, DNA damage and proliferation in multicellular tumor spheroids. BMC Cancer. 2017;17(1):338. doi: 10.1186/s12885-017-3319-0
  40. Riffle S, Hegde RS. Modeling tumor cell adaptations to hypoxia in multicellular tumor spheroids. J Exp Clin Cancer Res. 2017;36(1):102. doi: 10.1186/s13046-017-0570-9
  41. Ong SM, Zhao Z, Arooz T, et al. Engineering a scaffold-free 3D tumor model for in vitro drug penetration studies. Biomaterials. 2010;31(6):1180–1190. doi: 10.1016/j.biomaterials.2009.10.049
  42. Pokharel D, Wijesinghe P, Oenarto V, et al. Deciphering cell-to-cell communication in acquisition of cancer traits: extracellular membrane vesicles are regulators of tissue biomechanics. OMICS. 2016;20(8):462–469. doi: 10.1089/omi.2016.0072
  43. Conrad CB. Mechanical adaptability of ovarian cancer tumor spheroids [dissertation]. 2021. doi: 10.13016/yl6l-abpu
  44. Jaiswal D, Moscato Z, Tomizawa Y, et al. Elastography of multicellular spheroids using 3D light microscopy. Biomed Opt Express. 2019;10(5):2409–2418. doi: 10.1364/BOE.10.002409
  45. Hansen O, Magonov S, Su C, et al. Discerning biomolecular interactions using kelvin probe technology. Langmuir. 2003;19(18):7514–7520. doi: 10.1021/la034333w
  46. Sahin O, Magonov S, Su C, et al. An atomic force microscope tip designed to measure time-varying nanomechanical forces. Nat Nanotechnol. 2007;2(8):507–514. doi: 10.1038/nnano.2007.226
  47. Chen D, Gan H, Huang X, et al. Effects of peripheral blood mononuclear cells morphology on vascular calcification in uremic patients on maintenance hemodialysis. Ther Apher Dial. 2012;16(2):173–180. doi: 10.1111/j.1744-9987.2011.01044.x
  48. Jin H, Xing X, Zhao H, et al. Detection of erythrocytes influenced by aging and type 2 diabetes using atomic force microscope. Biochem Biophys Res Commun. 2010;391(4):1698–1702. doi: 10.1016/j.bbrc.2009.12.133
  49. Zhang Y, Zhang W, Wang S, et al. Detection of erythrocytes in patients with multiple myeloma using atomic force microscopy. Scanning. 2012;34(5):295–301. doi: 10.1002/sca.21008
  50. Zhang Y, Zhang W, Wang S, et al. Detection of erythrocytes in patient with iron deficiency anemia using atomic force microscopy. Scanning. 2012;34(4):215–220. doi: 10.1002/sca.20296
  51. Fang Y, Iu CY, Lui CN, et al. Investigating dynamic structural and mechanical changes of neuroblastoma cells associated with glutamate-mediated neurodegeneration. Sci Rep. 2014;4:7074. doi: 10.1038/srep07074
  52. Kim KS, Cho CH, Park EK, et al. AFM-detected apoptotic changes in morphology and biophysical property caused by paclitaxel in Ishikawa and HeLa cells. PLoS One. 2012;7(1):e30066. doi: 10.1371/journal.pone.0030066
  53. Chen M, Zeng J, Ruan W, et al. Examination of the relationship between viscoelastic properties and the invasion of ovarian cancer cells by atomic force microscopy. Beilstein J Nanotechnol. 2020;11:568–582. doi: 10.3762/bjnano.11.45
  54. Lian S, Shi R, Huang X, et al. Artesunate attenuates glioma proliferation, migration and invasion by affecting cellular mechanical properties. Oncol Rep. 2016;36(2):984–990. doi: 10.3892/or.2016.4847
  55. Zhang L, Yang F, Cai JY, et al. In-situ detection of resveratrol inhibition effect on epidermal growth factor receptor of living MCF-7 cells by atomic force microscopy. Biosens Bioelectron. 2014;56:271–277. doi: 10.1016/j.bios.2014.01.024
  56. Pi J, Li B, Tu L, et al. Investigation of quercetin-induced HepG2 cell apoptosis-associated cellular biophysical alterations by atomic force microscopy. Scanning. 2016;38(2):100–112. doi: 10.1002/sca.21245
  57. Wang J, Wan Z, Liu W, et al. Atomic force microscope study of tumor cell membranes following treatment with anti-cancer drugs. Biosens Bioelectron. 2009;25(4):721–727. doi: 10.1016/j.bios.2009.08.011
  58. Cambria E, Coughlin MF, Floryan MA, et al. Linking cell mechanical memory and cancer metastasis. Nat Rev Cancer. 2024;24(3):216–228. doi: 10.1038/s41568-023-00656-5
  59. Müller DJ, Helenius J, Alsteens D, Dufrêne YF. Force probing surfaces of living cells to molecular resolution. Nat Chem Biol. 2009;5(6):383–390. doi: 10.1038/nchembio.181
  60. Haase K, Pelling AE. Investigating cell mechanics with atomic force microscopy. J R Soc Interface. 2015;12(104):20140970. doi: 10.1098/rsif.2014.0970
  61. Müller DJ, Dufrêne YF. Atomic force microscopy as a multifunctional molecular toolbox in nanobiotechnology. Nat Nanotechnol. 2008;3(5):261–269. doi: 10.1038/nnano.2008.100
  62. Schillers H, Rianna C, Schäpe J, et al. Standardized nanomechanical atomic force microscopy procedure (snap) for measuring soft and biological samples. Sci Rep. 2017;7(1):5117. doi: 10.1038/s41598-017-05383-0
  63. Andolfi L, Greco SLM, Tierno D, et al. Planar AFM macro-probes to study the biomechanical properties of large cells and 3D cell spheroids. Acta Biomater. 2019;94:505–513. doi: 10.1016/j.actbio.2019.05.072
  64. Guillaume L, Rigal L, Fehrenbach J, et al. Characterization of the physical properties of tumor-derived spheroids reveals critical insights for pre-clinical studies. Sci Rep. 2019;9(1):6597. doi: 10.1038/s41598-019-43090-0
  65. Gonzalez-Rodriguez D, Guevorkian K, Douezan S, Brochard-Wyart F. Soft matter models of developing tissues and tumors. Science. 2012;338(6109):910–917. doi: 10.1126/science.1226418
  66. Dolega ME, Delarue M, Ingremeau F, et al. Cell-like pressure sensors reveal increase of mechanical stress towards the core of multicellular spheroids under compression. Nat Commun. 2017;8:14056. doi: 10.1038/ncomms14056
  67. Zhang W, Wang S, Lin M, et al. Advances in experimental approaches for investigating cell aggregate mechanics. Acta Mechanica Solida Sinica. 2012;25:473–482. doi: 10.1016/S0894-9166(12)60042-1
  68. Bhat S, Jun D, Paul B, Dahms T. In Viscoelasticity in biological systems: A special focus on microbes // InTech. 2012. doi: 10.5772/49980
  69. Sergunova V, Leesment S, Kozlov A, et al. Investigation of red blood cells by atomic force microscopy. Sensors (Basel). 2022;22(5):2055. doi: 10.3390/s22052055
  70. Stewart MP, Toyoda Y, Hyman AA, Müller DJ. Tracking mechanics and volume of globular cells with atomic force microscopy using a constant-height clamp. Nat Protoc. 2012;7(1):143–154. doi: 10.1038/nprot.2011.434
  71. Guimarães CF, Gasperini L, Marques AP, Reis RL. The stiffness of living tissues and its implications for tissue engineering. Nature Reviews Materials. 2020:5(5):351–370. doi: 10.1038/s41578-019-0169-1
  72. Fischer-Friedrich E, Hyman AA, Jülicher F, et al. Quantification of surface tension and internal pressure generated by single mitotic cells. Sci Rep. 2014;4:6213. doi: 10.1038/srep06213
  73. Krause M, Te Riet J, Wolf K. Probing the compressibility of tumor cell nuclei by combined atomic force-confocal microscopy. Phys Biol. 2013;10(6):065002. doi: 10.1088/1478-3975/10/6/065002
  74. Lam WA, Chaudhuri O, Crow A, et al. Mechanics and contraction dynamics of single platelets and implications for clot stiffening. Nat Mater. 2011;10(1):61–66. doi: 10.1038/nmat2903
  75. Stewart MP, Helenius J, Toyoda Y, et al. Hydrostatic pressure and the actomyosin cortex drive mitotic cell rounding. Nature. 2019;571(7764):E5. doi: 10.1038/s41586-019-1327-8
  76. Viljoen A, Mathelié-Guinlet M, Ray A, et al. Force spectroscopy of single cells using atomic force microscopy. Nat Rev Methods Primers. 2021;1(63):63. doi: 10.1038/s43586-021-00062-x
  77. Stewart MP, Hodel AW, Spielhofer A, et al. Wedged AFM-cantilevers for parallel plate cell mechanics. Methods. 2013;60(2):186–194. doi: 10.1016/j.ymeth.2013.02.015
  78. Stylianopoulos T, Martin JD, Chauhan VP, et al. Causes, consequences, and remedies for growth-induced solid stress in murine and human tumors. Proc Natl Acad Sci U S A. 2012;109(38):15101–15108. doi: 10.1073/pnas.1213353109
  79. Jaiswal D, Cowley N, Bian Z, et al. Stiffness analysis of 3D spheroids using microtweezers. PLoS One. 2017;12(11):e0188346. doi: 10.1371/journal.pone.0188346

Supplementary files

Supplementary Files
Action
1. JATS XML

Copyright (c) 2024 Eco-Vector

Creative Commons License
This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License.

СМИ зарегистрировано Федеральной службой по надзору в сфере связи, информационных технологий и массовых коммуникаций (Роскомнадзор).
Регистрационный номер и дата принятия решения о регистрации СМИ: 

This website uses cookies

You consent to our cookies if you continue to use our website.

About Cookies