(Translated by https://www.hiragana.jp/)
Microfluidic cell culture - Wikipedia Jump to content

Microfluidic cell culture

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Dutchy2020 (talk | contribs) at 15:06, 25 May 2020. The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Microfluidic cell culture integrates knowledge from biology, biochemistry, engineering, and physics to develop devices and techniques for culturing, maintaining, analyzing, and experimenting with cells at the microscale.[1][2] It merges microfluidics, a set of technologies used for the manipulation of small fluid volumes (μみゅーL, nL, pL) within artificially fabricated microsystems, and cell culture, which involves the maintenance and growth of cells in a controlled laboratory environment.[3][4] Microfluidics has been used for cell biology studies as the dimensions of the microfluidic channels are well suited for the physical scale of cells (in the order of magnitude of 10 micrometers).[2] For example, eukaryotic cells have linear dimensions between 10-100 μみゅーm which falls within the range of microfluidic dimensions.[4] A key component of microfluidic cell culture is being able to mimic the cell microenvironment which includes soluble factors that regulate cell structure, function, behavior, and growth.[2] Another important component for the devices is the ability to produce stable gradients that are present in vivo as these gradients play a significant role in understanding chemotactic, durotactic, and haptotactic effects on cells.[2]

Fabrication

Some considerations for microfluidic devices relating to cell culture include:

Fabrication material is crucial as not all polymers are biocompatible, with some materials such as PDMS causing undesirable adsorption or absorption of small molecules.[9][10] Additionally, uncured PDMS oligomers can leach into the cell culture media, which can harm the microenvironment.[9] As an alternative to commonly used PDMS, there have been advances in the use of thermoplastics (e.g., polystyrene) as a replacement material.[11][12]

Spatial organization of cells in microscale devices largely depends on the culture region geometry for cells to perform functions in vivo.[13][14] For example, long, narrow channels may be desired to culture neurons.[13] The perfusion system chosen might also affect the geometry chosen. For example, in a system that incorporates syringe pumps, channels for perfusion inlet, perfusion outlet, waste, and cell loading would need to be added for the cell culture maintenance.[15] Perfusion in microfluidic cell culture is important to enable long culture periods on-chip and cell differentiation.[16]

Other critical aspects for controlling the microenvironment include: cell seeding density, reduction of air bubbles as they can rupture cell membranes, evaporation of media due to an insufficiently humid environment, and cell culture maintenance (i.e. regular, timely media changes).[17][16][18]

Cell's health is defined as the collective equilibrium activities of essential and specialized cellular processes; while a cell stressor is defined as a stimulus that causes excursion from its equilibrium state. Hence, cell health may be perturbed within microsystems based on platform design or operating conditions. Exposure to stressors within microsystems can impact cells through direct and indirect ways. Therefore, it is important to design the microfluidics system for cell culture in a manner that minimizes cell stress situations. For example, by minimizing cell suspension, by avoiding abrupt geometries (which tend to favor bubble formation), designing higher and wider channels (to avoid shear stress), avoid thermosensitive hydrogels...[19]

Advantages

Some of the major advantages of microfluidic cell culture include reduced sample volumes (especially important when using primary cells, which are often limited) and the flexibility to customize and study multiple microenvironments within the same device.[3] A reduced cell population can also be used in a microscale system (e.g., a few hundred cells) in comparison to macroscale culture systems (which often require 105 – 107 cells); this can make studying certain cell-cell interactions more accessible.[10] These reduced cell numbers make studying non-dividing or slow dividing cells (e.g., stem cells) easier than traditional culture methods (e.g., flasks, petri dishes, or well plates) due to the smaller sample volumes.[10][20] Given the small dimensions in microfluidics, laminar flow can be achieved, allowing manipulations with the culture system to be done easily without affecting other culture chambers.[20] Laminar flow is also useful as is it mimics in vivo fluid dynamics more accurately, often making microscale culture more relevant than traditional culture methods.[21] Compartmentalized microfluidic cultures have also been combined with live cell calcium imaging, where depolarizing stimuli have been delivered to the peripheral terminals of neurons, and calcium responses recorded in the cell body.[22] This technique has demonstrated a stark difference in the sensitivity of the peripheral terminals compared to the neuronal cell body to certain stimuli such as protons.[22] This gives an excellent example as to why it is so important to study the peripheral terminals in isolation using microfluidic cell culture devices.

Culture Platforms

Traditional cell culture

Traditional two-dimensional (2D) cell culture is cell culture that takes place on a flat surface, e.g. the bottom of a well-plate, and is known as the conventional method.[23] While these platforms are useful for growing and passaging cells to be used in subsequent experiments, they are not ideal environments to monitor cell responses to stimuli as cells cannot freely move or perform functions as observed in vivo that are dependent on cell-extracellular matrix material interactions.[23] To address this issue many methods have been developed to create a three-dimensional (3D) native cell environment. One example of a 3D method is the hanging drop, where a droplet with growing cells is suspended and hangs downwards, which allows cells to grow around and atop of one another, forming a spheroid.[24] The hanging drop method has been used to culture tumor cells but is limited to the geometry of a sphere.[25] Since the advent of poly(dimethylsiloxane) (PDMS) microfluidic device fabrication through soft lithography[26] microfluidic devices have progressed and have proven to be very beneficial for mimicking a natural 3D environment for cell culture.[27]

Microfluidic cell culture

Microfluidic devices make possible the study of a single cell to a few hundred cells in a 3D environment. Comparatively, macroscopic 2D cultures have 104 to 107 cells on a flat surface.[28] Microfluidics also allow for chemical gradients, the continuous flow of fresh media, high through put testing, and direct output to analytical instruments.[28] Additionally, open microfluidic cell cultures such as "microcanals" allow for direct physical manipulation of cells with micropipettes.[29] Many microfluidic systems employ the use of hydrogels as the extracellular matrix (ECM) support which can be modulated for fiber thickness and pore size and have been demonstrated to allow the growth of cancer cells.[30] Gel free 3D cell cultures have been developed to allow cells to grow in either a cell dense environment or an ECM poor environment.[31] Although these devices have proven very useful, there are certain disadvantages such as cells sticking to the PDMS surface, small molecules diffusing into the PDMS, and unreacted PDMS polymers washing into cell culture media.[28]

The use of 3D cell cultures in microfluidic devices has led to a field of study called organ-on-a-chip. The first report of these types of microfluidic cultures was used to study the toxicity of naphthalene metabolites on the liver and lung (Viravaidya et al.). These devices can grow a stripped-down version of an organ-like system that can be used to understand many biological processes.[23] By adding an additional dimension, more advanced cell architectures can be achieved, and cell behavior is more representative of in vivo dynamics; cells can engage in enhanced communication with neighboring cells and cell-extracellular matrix interactions can be modeled.[23][32] In these devices, chambers or collagen layers containing different cell types can interact with one another for multiple days while various channels deliver nutrients to the cells.[23][33] An advantage of these devices is that tissue function can be characterized and observed under controlled conditions (e.g., effect of shear stress on cells, effect of cyclic strain or other forces) to better understand the overall function of the organ.[23][34] While these 3D models ofter better model organ function on a cellular level compared with 2D models, there are still challenges. Some of the challenges include: imaging of the cells, control of gradients in static models (i.e., without a perfusion system), and difficulty recreating vasculature.[34] Despite these challenges, 3D models are still used as tools for studying and testing drug responses in pharmacological studies.[23] In recent years, there are microfluidic devices reproducing the complex in vivo microvascular network. Organs-on-a-chip have also been used to replicate very complex systems like lung epithelial cells in an exposed airway and provides valuable insight for how multicellular systems and tissues function in vivo.[35] These devices are able to create a physiologically realistic 3D environment, which is desirable as a tool for drug screening, drug delivery, cell-cell interactions, tumor metastasis etc.[36][37] In one study, researchers grew tumor cells and tested the drug delivery of cis platin, resveratrol, tirapazamine (TPZ) and then measured the effects the drugs have on cell viability.[38]

See also

References

  1. ^ a b Bhatia SN, Ingber DE (August 2014). "Microfluidic organs-on-chips". Nature Biotechnology. 32 (8): 760–72. doi:10.1038/nbt.2989. PMID 25093883.
  2. ^ a b c d Young EW, Beebe DJ (March 2010). "Fundamentals of microfluidic cell culture in controlled microenvironments". Chemical Society Reviews. 39 (3): 1036–48. doi:10.1039/b909900j. PMC 2967183. PMID 20179823.
  3. ^ a b Mehling M, Tay S (February 2014). "Microfluidic cell culture". Current Opinion in Biotechnology. 25: 95–102. doi:10.1016/j.copbio.2013.10.005. PMID 24484886.
  4. ^ a b Whitesides GM (July 2006). "The origins and the future of microfluidics". Nature. 442 (7101): 368–73. Bibcode:2006Natur.442..368W. doi:10.1038/nature05058. PMID 16871203.
  5. ^ Cho, Brenda S.; Schuster, Timothy G.; Zhu, Xiaoyue; Chang, David; Smith, Gary D.; Takayama, Shuichi (2003-04-01). "Passively Driven Integrated Microfluidic System for Separation of Motile Sperm". Analytical Chemistry. 75 (7): 1671–1675. doi:10.1021/ac020579e. ISSN 0003-2700. PMID 12705601.
  6. ^ Zimmermann M, Schmid H, Hunziker P, Delamarche E (January 2007). "Capillary pumps for autonomous capillary systems". Lab on a Chip. 7 (1): 119–25. doi:10.1039/b609813d. PMID 17180214.
  7. ^ Walker G, Beebe DJ (August 2002). "A passive pumping method for microfluidic devices". Lab on a Chip. 2 (3): 131–4. CiteSeerX 10.1.1.118.5648. doi:10.1039/b204381e. PMID 15100822.
  8. ^ Kim L, Toh YC, Voldman J, Yu H (June 2007). "A practical guide to microfluidic perfusion culture of adherent mammalian cells". Lab on a Chip. 7 (6): 681–94. doi:10.1039/b704602b. PMID 17538709.
  9. ^ a b Regehr KJ, Domenech M, Koepsel JT, Carver KC, Ellison-Zelski SJ, Murphy WL, Schuler LA, Alarid ET, Beebe DJ (August 2009). "Biological implications of polydimethylsiloxane-based microfluidic cell culture". Lab on a Chip. 9 (15): 2132–9. doi:10.1039/b903043c. PMC 2792742. PMID 19606288.
  10. ^ a b c Halldorsson S, Lucumi E, Gómez-Sjöberg R, Fleming RM (January 2015). "Advantages and challenges of microfluidic cell culture in polydimethylsiloxane devices". Biosensors & Bioelectronics. 63: 218–231. doi:10.1016/j.bios.2014.07.029. PMID 25105943.
  11. ^ Berthier E, Young EW, Beebe D (April 2012). "Engineers are from PDMS-land, Biologists are from Polystyrenia". Lab on a Chip. 12 (7): 1224–37. doi:10.1039/c2lc20982a. PMID 22318426.
  12. ^ van Midwoud PM, Janse A, Merema MT, Groothuis GM, Verpoorte E (May 2012). "Comparison of biocompatibility and adsorption properties of different plastics for advanced microfluidic cell and tissue culture models". Analytical Chemistry. 84 (9): 3938–44. doi:10.1021/ac300771z. PMID 22444457.
  13. ^ a b Rhee SW, Taylor AM, Tu CH, Cribbs DH, Cotman CW, Jeon NL (January 2005). "Patterned cell culture inside microfluidic devices". Lab on a Chip. 5 (1): 102–7. doi:10.1039/b403091e. hdl:10371/7982. PMID 15616747.
  14. ^ Folch A, Toner M (1998-01-01). "Cellular micropatterns on biocompatible materials". Biotechnology Progress. 14 (3): 388–92. doi:10.1021/bp980037b. PMID 9622519.
  15. ^ Hung PJ, Lee PJ, Sabounchi P, Lin R, Lee LP (January 2005). "Continuous perfusion microfluidic cell culture array for high-throughput cell-based assays". Biotechnology and Bioengineering. 89 (1): 1–8. doi:10.1002/bit.20289. PMID 15580587.
  16. ^ a b Tourovskaia A, Figueroa-Masot X, Folch A (January 2005). "Differentiation-on-a-chip: a microfluidic platform for long-term cell culture studies". Lab on a Chip. 5 (1): 14–9. doi:10.1039/b405719h. PMID 15616734.
  17. ^ Meyvantsson I, Beebe DJ (2008-06-13). "Cell culture models in microfluidic systems". Annual Review of Analytical Chemistry. 1 (1): 423–49. Bibcode:2008ARAC....1..423M. doi:10.1146/annurev.anchem.1.031207.113042. PMID 20636085.
  18. ^ Yu H, Alexander CM, Beebe DJ (June 2007). "Understanding microchannel culture: parameters involved in soluble factor signaling". Lab on a Chip. 7 (6): 726–30. doi:10.1039/b618793e. PMID 17538714.
  19. ^ Varma, Sarvesh; Voldman, Joel (2018). "Caring for cells in microsystems: principles and practices of cell-safe device design and operation". Lab on a Chip. 18 (22): 3333–3352. doi:10.1039/C8LC00746B. ISSN 1473-0197. PMC 6254237. PMID 30324208.
  20. ^ a b Gómez-Sjöberg R, Leyrat AA, Pirone DM, Chen CS, Quake SR (November 2007). "Versatile, fully automated, microfluidic cell culture system". Analytical Chemistry. 79 (22): 8557–63. doi:10.1021/ac071311w. PMID 17953452.
  21. ^ Cimetta E, Vunjak-Novakovic G (September 2014). "Microscale technologies for regulating human stem cell differentiation". Experimental Biology and Medicine. 239 (9): 1255–63. doi:10.1177/1535370214530369. PMC 4476254. PMID 24737735.
  22. ^ a b Clark, Alex J.; Menendez, Guillermo; AlQatari, Mona; Patel, Niral; Arstad, Erik; Schiavo, Giampietro; Koltzenburg, Martin (2018). "Functional imaging in microfluidic chambers reveals sensory neuron sensitivity is differentially regulated between neuronal regions". PAIN. 159 (7): 1413–1425. doi:10.1097/j.pain.0000000000001145. ISSN 0304-3959. PMID 29419650.
  23. ^ a b c d e f g Bhatia SN, Ingber DE (August 2014). "Microfluidic organs-on-chips". Nature Biotechnology. 32 (8): 760–72. doi:10.1038/nbt.2989. PMID 25093883.
  24. ^ Keller, Gordon M (January 1995). "In vitro differentiation of embryonic stem cells". Current Opinion in Cell Biology. 7 (6): 862–869. doi:10.1016/0955-0674(95)80071-9. PMID 8608017.
  25. ^ Kelm, Jens M.; Timmins, Nicholas E.; Brown, Catherine J.; Fussenegger, Martin; Nielsen, Lars K. (2003-05-21). "Method for generation of homogeneous multicellular tumor spheroids applicable to a wide variety of cell types". Biotechnology and Bioengineering. 83 (2): 173–180. doi:10.1002/bit.10655. ISSN 0006-3592. PMID 12768623.
  26. ^ Duffy, David C.; McDonald, J. Cooper; Schueller, Olivier J. A.; Whitesides, George M. (December 1998). "Rapid Prototyping of Microfluidic Systems in Poly(dimethylsiloxane)". Analytical Chemistry. 70 (23): 4974–4984. doi:10.1021/ac980656z. ISSN 0003-2700. PMID 21644679.
  27. ^ Gupta, Nilesh; Liu, Jeffrey R.; Patel, Brijeshkumar; Solomon, Deepak E.; Vaidya, Bhuvaneshwar; Gupta, Vivek (March 2016). "Microfluidics-based 3D cell culture models: Utility in novel drug discovery and delivery research: Gupta et al". Bioengineering & Translational Medicine. 1 (1): 63–81. doi:10.1002/btm2.10013. PMC 5689508. PMID 29313007.
  28. ^ a b c Halldorsson S, Lucumi E, Gómez-Sjöberg R, Fleming RM (January 2015). "Advantages and challenges of microfluidic cell culture in polydimethylsiloxane devices". Biosensors & Bioelectronics. 63: 218–231. doi:10.1016/j.bios.2014.07.029. PMID 25105943.
  29. ^ Hsu, Chia-Hsien; Chen, Chihchen; Folch, Albert (2004-10-07). ""Microcanals" for micropipette access to single cells in microfluidic environments". Lab on a Chip. 4 (5): 420–424. doi:10.1039/B404956J. ISSN 1473-0189.
  30. ^ Ma, Yu-Heng Vivian; Middleton, Kevin; You, Lidan; Sun, Yu (2018-04-09). "A review of microfluidic approaches for investigating cancer extravasation during metastasis". Microsystems & Nanoengineering. 4: 17104. doi:10.1038/micronano.2017.104. ISSN 2055-7434.
  31. ^ Ong, Siew-Min; Zhang, Chi; Toh, Yi-Chin; Kim, So Hyun; Foo, Hsien Loong; Tan, Choon Hong; van Noort, Danny; Park, Sungsu; Yu, Hanry (August 2008). "A gel-free 3D microfluidic cell culture system". Biomaterials. 29 (22): 3237–3244. doi:10.1016/j.biomaterials.2008.04.022. PMID 18455231.
  32. ^ Pampaloni F, Reynaud EG, Stelzer EH (October 2007). "The third dimension bridges the gap between cell culture and live tissue". Nature Reviews. Molecular Cell Biology. 8 (10): 839–45. doi:10.1038/nrm2236. PMID 17684528.
  33. ^ Huh D, Hamilton GA, Ingber DE (December 2011). "From 3D cell culture to organs-on-chips". Trends in Cell Biology. 21 (12): 745–54. doi:10.1016/j.tcb.2011.09.005. PMC 4386065. PMID 22033488.
  34. ^ a b van Duinen V, Trietsch SJ, Joore J, Vulto P, Hankemeier T (December 2015). "Microfluidic 3D cell culture: from tools to tissue models". Current Opinion in Biotechnology. 35: 118–26. doi:10.1016/j.copbio.2015.05.002. PMID 26094109.
  35. ^ Benam, Kambez H; Villenave, Remi; Lucchesi, Carolina; Varone, Antonio; Hubeau, Cedric; Lee, Hyun-Hee; Alves, Stephen E; Salmon, Michael; Ferrante, Thomas C (February 2016). "Small airway-on-a-chip enables analysis of human lung inflammation and drug responses in vitro". Nature Methods. 13 (2): 151–157. doi:10.1038/nmeth.3697. ISSN 1548-7091. PMID 26689262.
  36. ^ Soroush F, Zhang T, King DJ, Tang Y, Deosarkar S, Prabhakarpandian B, Kilpatrick LE, Kiani MF (November 2016). "A novel microfluidic assay reveals a key role for protein kinase C δでるた in regulating human neutrophil-endothelium interaction". Journal of Leukocyte Biology. 100 (5): 1027–1035. doi:10.1189/jlb.3ma0216-087r. PMC 5069089. PMID 27190303.
  37. ^ Tang Y, Soroush F, Deosarkar S, Wang B, Pandian P, Kiani MF (April 2016). "A Novel Synthetic Tumor Platform for Screening Drug Delivery systems". The FASEB Journal. doi:10.1096/fasebj.30.1_supplement.698.7 (inactive 2020-01-22).{{cite journal}}: CS1 maint: DOI inactive as of January 2020 (link) CS1 maint: unflagged free DOI (link)
  38. ^ Patra, Bishnubrata; Peng, Chien-Chung; Liao, Wei-Hao; Lee, Chau-Hwang; Tung, Yi-Chung (August 2016). "Drug testing and flow cytometry analysis on a large number of uniform sized tumor spheroids using a microfluidic device". Scientific Reports. 6 (1): 21061. Bibcode:2016NatSR...621061P. doi:10.1038/srep21061. ISSN 2045-2322. PMC 4753452. PMID 26877244.