Physics:Cell culturing in open microfluidics

From HandWiki
The printable version is no longer supported and may have rendering errors. Please update your browser bookmarks and please use the default browser print function instead.

}}

Open microfluidics can be employed in the multidimensional culturing of cell types for various applications including organ-on-a-chip studies, oxygen-driven reactions, neurodegeneration, cell migration, and other cellular pathways.[1][2]

Usage and benefits

The use of conventional microfluidic devices for cell studies has already improved upon the cost effectiveness and sample volume requirement, however using open microfluidic channels adds the benefit of removing syringe pumps to drive flow, now governed by surface tensions that drive spontaneous capillary flow (SCF), and exposes cells to the surrounding environment.[3][4][5] The miniaturization of this process allows for improved sensitivity, high throughput, and ease of manipulation and integration, as well as dimensions that can be more physiologically relevant.[4][5][6][7][1] The benefits of both open and closed microfluidic platforms have allowed the option for the combination of the two, where the device is open for the introduction and culturing of cells, and can be sealed prior to analysis.[3]

Design

Cells and proteins can be patterned in microfluidic devices with one of the channel walls exposed in different geometries and designs depending on the behaviors and interactions to be studied, such as quorum sensing or co-culturing of several types of cells.[6][8] A majority of cell culturing has been carried out by introducing the cells in a perfused conditioned medium to simulate the desired cell populations in traditional close-channel microfluidic devices. The challenge to support the cell growth and simultaneously study multiple cell types in a single device with an exposed channel is that the interactions between cells in this medium needs to be controlled since the timing and location of the interactions is critical.[9] This issue can be addressed in several ways including the modification of the device design, using droplet microfluidics, and cell sorting.[9][10] Not only does this allow for the ease of manipulating the environment of the cells, but having an open channel wall allows for a better understanding of biological interactions at this interface.[9] Creating designs of microfluidic platforms with different compartments that are isolated and have different dimensions allows for co-culturing of several types of cells.[6] These devices often incorporate droplet formation to encapsulate cells and act as transport and reaction vehicles in two or more immiscible phases, making it possible to carry out numerous parallel analyses using different conditions.[5][11] Open microfluidics has also been coupled with fluorescence-activated cell sorting (FACS) to allow for cells to be contained in individually sorted compartments in an open microfluidic network for culturing in an exposed environment.[10] The exposure of one of the channel walls introduces the issue of evaporation and therefore cell loss, however this issue can be minimized by using droplet microfluidics where the cell-containing droplets are submerged in a fluorinated oil.[12] Although evaporation is a major disadvantage of using an open microfluidic system for cell culturing, the advantages over a closed system include ease of manipulation and access to the cells. For certain applications, such as the study of drug transport and lung function using alveolar epithelium cells, air exposure to is essential for developing the lungs.[7]

PDMS

Polydimethylsiloxane (PDMS) is a common material for open microfluidic devices that introduces additional advantages and disadvantages. The adsorption of small biological molecules from cell culturing samples as well as the release of oligomers into the culture medium have both been posed as issues of using PDMS for biological studies, however these can be reduced by adopting pretreatment procedures to create optimal environments.[13] Advantages of using PDMS include the ease of surface modification, low cost, biocompatibility, and optical transparency.[14] In addition, PDMS is an attractive material to use for generating oxygen gradients for cell culturing in studies that involve monitoring ROS governed cellular pathways due to its oxygen permeability.[15] Plastics such as polystyrene can be used to create microfluidic devices by embossing and bonding methods, CNC milling, injection molding, or stereolithography.[16][17] Devices created with polystyrene by these methods include microfluidic platforms that integrate several microfluidic systems, creating arrays to study several cell cultures simultaneously.[16] Another type of material that is used for open-microfluidic cell culturing is paper-based microfluidics. Cell culturing on paper-based microfluidic devices is accomplished either by encapsulating cells in a hydrogel or directly seeding them in stacked cellulose filter papers and the cell culture medium is passively transported to the culture areas.[18] A major advantage of this type of open-microfluidics includes the low cost, the variety of dimensions of porous papers that are commercially available, improved cell viability, adhesion, and migration over tissue culture plates.[18][19] In addition, it is an attractive substrate for 3D cell culture devices due to its ability to incorporate essential characteristics such as oxygen and nutrient gradients, fluid flow that can control cell migration, and stacking filter papers with different cells suspended in hydrogel to monitor cellular interactions or complex populations.[19][20][21]

References

  1. 1.0 1.1 Lin, Dongguo; Li, Peiwen; Lin, Jinqiong; Shu, Bowen; Wang, Weixin; Zhang, Qiong; Yang, Na; Liu, Dayu et al. (2017-10-31). "Orthogonal Screening of Anticancer Drugs Using an Open-Access Microfluidic Tissue Array System". Analytical Chemistry 89 (22): 11976–11984. doi:10.1021/acs.analchem.7b02021. ISSN 0003-2700. PMID 29053257. 
  2. Malboubi, Majid; Jayo, Asier; Parsons, Maddy; Charras, Guillaume (August 2015). "An open access microfluidic device for the study of the physical limits of cancer cell deformation during migration in confined environments". Microelectronic Engineering 144: 42–45. doi:10.1016/j.mee.2015.02.022. ISSN 0167-9317. PMID 26412914. 
  3. 3.0 3.1 Lovchik, Robert D.; Bianco, Fabio; Tonna, Noemi; Ruiz, Ana; Matteoli, Michela; Delamarche, Emmanuel (May 2010). "Overflow Microfluidic Networks for Open and Closed Cell Cultures on Chip". Analytical Chemistry 82 (9): 3936–3942. doi:10.1021/ac100771r. ISSN 0003-2700. PMID 20392062. 
  4. 4.0 4.1 Lee, J. J., Berthier, J., Brakke, K. A., Dostie, A. M., Theberge, A. B., & Berthier, E. (2018). Droplet Behavior in Open Biphasic Microfluidics. Langmuir, 34(18), 5358–5366. doi:10.1021/acs.langmuir.8b00380
  5. 5.0 5.1 5.2 Schneider, Thomas; Kreutz, Jason; Chiu, Daniel T. (2013-03-15). "The Potential Impact of Droplet Microfluidics in Biology". Analytical Chemistry 85 (7): 3476–3482. doi:10.1021/ac400257c. ISSN 0003-2700. PMID 23495853. 
  6. 6.0 6.1 6.2 Lee, Sung Hoon; Heinz, Austen James; Shin, Sunghwan; Jung, Yong-Gyun; Choi, Sung-Eun; Park, Wook; Roe, Jung-Hye; Kwon, Sunghoon (April 2010). "Capillary Based Patterning of Cellular Communities in Laterally Open Channels". Analytical Chemistry 82 (7): 2900–2906. doi:10.1021/ac902903q. ISSN 0003-2700. PMID 20210331. https://figshare.com/articles/Capillary_Based_Patterning_of_Cellular_Communities_in_Laterally_Open_Channels/2780779. 
  7. 7.0 7.1 Nalayanda, Divya D.; Puleo, Christopher; Fulton, William B.; Sharpe, Leilani M.; Wang, Tza-Huei; Abdullah, Fizan (2009-05-30). "An open-access microfluidic model for lung-specific functional studies at an air-liquid interface". Biomedical Microdevices 11 (5): 1081–1089. doi:10.1007/s10544-009-9325-5. ISSN 1387-2176. PMID 19484389. 
  8. Boedicker, James Q.; Vincent, Meghan E.; Ismagilov, Rustem F. (2009-07-27). "Microfluidic Confinement of Single Cells of Bacteria in Small Volumes Initiates High-Density Behavior of Quorum Sensing and Growth and Reveals Its Variability". Angewandte Chemie International Edition 48 (32): 5908–5911. doi:10.1002/anie.200901550. ISSN 1433-7851. PMID 19565587. 
  9. 9.0 9.1 9.2 Kaigala, G. V., Lovchik, R. D., & Delamarche, E. (2012). Microfluidics in the “open Space” for performing localized chemistry on biological interfaces. Angewandte Chemie - International Edition, 51(45), 11224–11240. doi:10.1002/anie.201201798
  10. 10.0 10.1 Birchler, Axel; Berger, Mischa; Jäggin, Verena; Lopes, Telma; Etzrodt, Martin; Misun, Patrick Mark; Pena-Francesch, Maria; Schroeder, Timm et al. (2016-01-06). "Seamless Combination of Fluorescence-Activated Cell Sorting and Hanging-Drop Networks for Individual Handling and Culturing of Stem Cells and Microtissue Spheroids". Analytical Chemistry 88 (2): 1222–1229. doi:10.1021/acs.analchem.5b03513. ISSN 0003-2700. PMID 26694967. 
  11. Casavant, B. P., Berthier, E., Theberge, A. B., Berthier, J., Montanez-Sauri, S. I., Bischel, L. L., … Beebe, D. J. (2013). Suspended microfluidics. Proceedings of the National Academy of Sciences, 110(25), 10111–10116. doi:10.1073/pnas.1302566110
  12. Li, Chao; Yu, Jiaquan; Schehr, Jennifer; Berry, Scott M.; Leal, Ticiana A.; Lang, Joshua M.; Beebe, David J. (2018-05-08). "Exclusive Liquid Repellency: An Open Multi-Liquid-Phase Technology for Rare Cell Culture and Single-Cell Processing". ACS Applied Materials & Interfaces 10 (20): 17065–17070. doi:10.1021/acsami.8b03627. ISSN 1944-8244. PMID 29738227. 
  13. Regehr, Keil J.; Domenech, Maribella; Koepsel, Justin T.; Carver, Kristopher C.; Ellison-Zelski, Stephanie J.; Murphy, William L.; Schuler, Linda A.; Alarid, Elaine T. et al. (2009). "Biological implications of polydimethylsiloxane-based microfluidic cell culture". Lab on a Chip 9 (15): 2132–2139. doi:10.1039/b903043c. ISSN 1473-0197. PMID 19606288. 
  14. Halldorsson, S., Lucumi, E., Gómez-Sjöberg, R., & Fleming, R. M. T. (2015). Advantages and challenges of microfluidic cell culture in polydimethylsiloxane devices. Biosensors and Bioelectronics, 63, 218–231. doi:10.1016/j.bios.2014.07.029
  15. Lo, J. F., Sinkala, E., & Eddington, D. T. (2010). Oxygen gradients for open well cellular cultures via microfluidic substrates. Lab on a Chip, 10(18), 2394–2401. doi:10.1039/c004660d
  16. 16.0 16.1 Young, Edmond W. K.; Berthier, Erwin; Guckenberger, David J.; Sackmann, Eric; Lamers, Casey; Meyvantsson, Ivar; Huttenlocher, Anna; Beebe, David J. (2011-02-15). "Rapid Prototyping of Arrayed Microfluidic Systems in Polystyrene for Cell-Based Assays". Analytical Chemistry 83 (4): 1408–1417. doi:10.1021/ac102897h. ISSN 0003-2700. PMID 21261280. 
  17. Guckenberger, David J.; de Groot, Theodorus E.; Wan, Alwin M. D.; Beebe, David J.; Young, Edmond W. K. (2015). "Micromilling: a method for ultra-rapid prototyping of plastic microfluidic devices". Lab on a Chip 15 (11): 2364–2378. doi:10.1039/C5LC00234F. ISSN 1473-0197. PMID 25906246. 
  18. 18.0 18.1 Tao, F. F., Xiao, X., Lei, K. F., & Lee, I. C. (2015). Paper-based cell culture microfluidic system. Biochip Journal, 9(2), 97–104. doi:10.1007/s13206-015-9202-7
  19. 19.0 19.1 Ng, K., Gao, B., Yong, K. W., Li, Y., Shi, M., Zhao, X., … Xu, F. (2017). Paper-based cell culture platform and its emerging biomedical applications. Materials Today, 20(1), 32–44. doi:10.1016/j.mattod.2016.07.001
  20. Mosadegh, Bobak; Dabiri, Borna E.; Lockett, Matthew R.; Derda, Ratmir; Campbell, Patrick; Parker, Kevin Kit; Whitesides, George M. (2014-02-12). "Three-Dimensional Paper-Based Model for Cardiac Ischemia". Advanced Healthcare Materials 3 (7): 1036–1043. doi:10.1002/adhm.201300575. ISSN 2192-2640. PMID 24574054. 
  21. YAN, Wei; ZHANG, Qiong; CHEN, Bin; LIANG, Guang-Tie; LI, Wei-Xuan; ZHOU, Xiao-Mian; LIU, Da-Yu (June 2013). "Study on Microenvironment Acidification by Microfluidic Chip with Multilayer-paper Supported Breast Cancer Tissue". Chinese Journal of Analytical Chemistry 41 (6): 822–827. doi:10.1016/s1872-2040(13)60661-1. ISSN 1872-2040.