Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 1402 word(s) 1402 2021-12-28 04:57:59 |
2 update layout and reference -11 word(s) 1391 2022-01-12 03:49:02 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Ribeiro, J.; Lima, R. Applications of polydimethylsiloxane (PDMS) in Engineering. Encyclopedia. Available online: https://encyclopedia.pub/entry/18079 (accessed on 01 May 2024).
Ribeiro J, Lima R. Applications of polydimethylsiloxane (PDMS) in Engineering. Encyclopedia. Available at: https://encyclopedia.pub/entry/18079. Accessed May 01, 2024.
Ribeiro, João, Rui Lima. "Applications of polydimethylsiloxane (PDMS) in Engineering" Encyclopedia, https://encyclopedia.pub/entry/18079 (accessed May 01, 2024).
Ribeiro, J., & Lima, R. (2022, January 11). Applications of polydimethylsiloxane (PDMS) in Engineering. In Encyclopedia. https://encyclopedia.pub/entry/18079
Ribeiro, João and Rui Lima. "Applications of polydimethylsiloxane (PDMS) in Engineering." Encyclopedia. Web. 11 January, 2022.
Applications of polydimethylsiloxane (PDMS) in Engineering
Edit

Polydimethylsiloxane (PDMS) is an elastomer with excellent optical, electrical and mechanical properties, which makes it well-suited for several engineering applications. Due to its biocompatibility, PDMS is widely used for biomedical purposes. Some properties can be improved by adding additives.

polydimethylsiloxane PDMS properties PDMS applications PDMS PDMS composites mechanical properties optical properties biocompatibility

1. Introduction

Polydimethylsiloxane (PDMS) is an elastomeric polymer with interesting properties for biomedical applications, including physiological indifference, excellent resistance to biodegradation, biocompatibility, chemical stability, gas permeability, good mechanical properties, excellent optical transparency and simple fabrication by replica moulding [1][2][3][4][5]. Due to these characteristics, PDMS has been widely used in micropumps [6], catheter surfaces [7], dressings and bandages [8], microvalves [9], optical systems [10][11], in the in vitro study of diseases [12][13], in implants [14][15], in microfluidics and photonics [16][17][18][19]. Moreover, soft-lithography technology has driven the use of PDMS in microelectromechanical systems (MEMS) applications and in microfluidic components [17][18][20]. Soft-lithography techniques such as micro-contact printing, replica moulding, micro-transfer moulding, micro-moulding in capillaries and solvent-assisted micro-moulding usually require the use of PDMS to create an elastomeric stamp or mould that incorporates nano- and microstructures for the transfer of patterns onto a subsequent substrate [18][21].
MEMS are approaches that use electronic and mechanical technologies to deal with biomedical problems on the micro-scale [22]. MEMS-based devices have been widely used in the biomedical area for applications such as diagnostics and therapeutics. These systems can be microsensors or microtransducers, and are helpful in areas such as physics, mechanics, electronics and biomedicals, as they can provide very precise and fast results [23]. The investigation and improvement of already existing MEMS are more and more common. As they are increasingly commercialized, the necessity to find processes and materials that enable mass production while reducing cost has emerged [21]. MEMS are traditionally silicon-based and the pursuit for a more biologically friendly material is needed. Polymers allow rapid prototyping and mass production techniques as well as having a lower cost in relation to silicon, making them particularly attractive for the development of MEMS [21]. Photolithography is the most commonly used technique in microfabrication, however, this method is expensive [24]. With the introduction of polymers in microsystems, new manufacturing techniques have been studied, such as soft-lithography, which can be a cheaper method comparatively to photolithography, even when a costly mould is needed for patterning; once a mould is created, it can be reused several times [20]. Additionally, there are alternatives which are attempting to reduce the cost of the moulds, relying on cleanroom less approaches [25]. Candidate polymers for the production of MEMS are polycarbonate (PC), polymethylmethacrylate (PMMA), polyvinylchloride (PVC), polyethylene (PE) and PDMS [21].
Additionally, PDMS is the most commonly used material in the manufacturing of microfluidic devices, which are an important technology for the development of systems such as drug delivery, DNA sequencing, clinical diagnostics, point of care testing and chemical synthesis [26]. The used materials in these systems should be biocompatible, optically transparent and provide fast prototyping and low fabrication cost [27], features found in PDMS.
In addition to applications in microfluidics, PDMS has been widely used in the fabrication of biomodels (flow phantom) for the in vitro hemodynamic study of diseases such as aneurysms and stenosis [28][29][30][31]. The biomodels developed in PDMS allow good replicability of the lumen of the arteries and good transparency, being ideal for the application of optical techniques of micro particle image velocimetry (micro-PIV), particle image velocimetry (PIV), particle tracking velocimetry (PTV) and non-evasive techniques [32][33][34]. These experimental tests have provided a greater understanding of these pathologies, validated numerical techniques and tested medical devices such as stents [35][36][37].
PDMS has also been investigated in the field of medical implants [38][39][40][41][42]. These types of implants are usually made with titanium or its alloys; however, such materials do not allow good osseointegration [39]. In order to overcome this limitation, PDMS has been studied to produce coatings with microscale features that help the bonding between the implant and the bone. The main characteristics for its use in implants are its high biocompatibility, excellent resistance to biodegradation and flexibility, which makes PDMS one of the most successful polymers in implanted devices, presenting only mild foreign body reactions [43][44][45]. Common applications include cardiac pacemakers, cuff and book electrodes in the PNS, cochlear implants, bladder and pain controllers and planar electrode arrays in the CNS [45][46].

2. PDMS Properties

Silicon, glass and polymers are the typical materials used for micro devices fabrication: silicon, because of its thermal conductivity and the availability of advanced fabrication technologies; glass, mainly due to its transparency; polymers, because of its low cost, optical transparency and flexibility. Compared to glass and silicon, PDMS turns out to be the most promising elastomer, because the other two materials have a high manufacturing cost, require greater labour intensity and are rigid in nature. The variable elasticity of PDMS in medical applications is also favourable; its modulus of elasticity is 1–3 MPa (compared to ~50 GPa of glass) [2][47]. PDMS is also chemically inert, thermally stable, permeable to gases, simple to handle and manipulate, exhibits isotropic and homogeneous properties and can replicate submicron features to develop microstructures [19][21][48]. Additionally, this elastomer is optically transparent, can work as a thermal and electrical insulator and degrades quickly in the natural environment [49]. PDMS presents a hyperelastic behaviour, which is the ability of a material to undergo large deformations before rupture [50]. This characteristic is also found in biological tissues and, for that reason, PDMS is a well-suited material to mimic, for example, blood vessels [49]. Another characteristic of this elastomer is its biocompatibility, which means that PDMS is compatible with biologic tissues [49]. PDMS presents a transmittance up to 90% for the wavelength from 390 nm to 780 nm [51][52][53] and, due to this characteristic, PDMS-based microsystems allow the direct observation of the mimicked blood flow inside the mimicked vessels and the integration of optical detection systems, hence playing an important role in this field.
With the purpose of extending the lifespan of a chip, PDMS is used to embed or encapsulate electronic components by casting. Due to its thermal and electrical insulation capability, PDMS protects the components from environmental factors and mechanical shock within a large temperature range (−50–200 °C) [23][48]. In Table 1, some physical properties of PDMS are listed.
Table 1. Typical properties of cured PDMS.
Property (Unity) Result References
Transmittance at range 390 nm to 780 nm (%) 75–92 [54][55]
Index of refraction 1.4 [56]
Thermal conductivity (W/m∙K) 0.2–0.27 [57][58]
Specific heat (kJ/kg∙K) 1.46 [56]
Dielectric strength (kV/mm) 19 [57]
Dielectric constant 2.3–2.8 [56]
Electrical conductivity (ohm∙m) 4 × 1013 [56]
Volume resistivity (ohm∙cm) 2.9 × 1014 [57]
Young’s modulus (kPa) 360–870 [59]
Poisson ratio 0.5 [60]
Tensile strength (MPa) 2.24–6.7 [56][57]
Hardness (Shore A) 41–43 [55][61]
Viscosity (Pa∙s) 3.5 [57]
Hydrophobicity—contact angle (°) ~108° ± 7° [62]
Melting Point (°C) −49.9 to −40 [63]
Despite these advantages, PDMS has some properties that can present a limitation in some applications. Due to its CH3 groups, PDMS presents a hydrophobic surface (contact angle with water ~108° ± 7°) [62][64][65], often limiting its application in solutions composed of biological samples [66]. Additionally, PDMS tends to swell when combined with certain reagents [17][48]. In some applications, the absorption of small molecules flowing through the channels makes it difficult to quantitatively analyse experiments in proteomic drug discovery and cell culture [67][68]. In microchannels, the hydrophobicity of PDMS generates complications that include impedance to the flow of polar liquids, which makes it difficult to wet its surface with aqueous solvents [49]. On the other hand, much effort has been made to make the PDMS surface hydrophilic and resistant to protein adsorption [19][69][70][71][72][73].
Strategies employed in attempting to solve PDMS hydrophobicity include surface activation methods such as: oxygen plasma; UV/ozone treatments; corona discharges, which are widely used for PDMS surface oxidation to promote microchannel wettability. The main benefits of these methods are the short treatment time and easy operation; however, the PDMS surface recovers its hydrophobicity when in contact with air within a few minutes [74][75][76].
Another method is physisorption, which is a simple and efficient approach that relies on surface hydrophobic or electrostatic interactions. This method includes the following techniques: layer-by-layer deposition; non-ionic surfactants; charged polymers. The disadvantages are the lack of covalent bonds between PDMS and surface modifiers, which lead to the loss of modifiers quickly through desorption [77][78][79].
In order to improve the difficulties encountered in physisorption, chemical modification methods allow for maintaining a long-term stability of the modified surface. These methods include: chemical vapor deposition, surface segregation and self-assembled monolayers, silanization, and polymer brushes via grafting methods [1][62][80][81][82].
Adding waxes such as paraffin or beeswax to PDMS has been demonstrated to potentially increase the corrosion resistance, hydrophobicity and thermal and optical properties of PDMS, which is useful in applications such as sensors, wearable devices and superhydrophobic coating [83].
This entry is adapted from 10.3390/jfb13010002 and 10.3390/polym13234258

References

  1. Poll, M.; Zhou, F.; Ramstedt, M.; Hu, L.; Huck, W. A Self-assembly approach to chemical micropatterning of Poly(dimethylsiloxane). Angew. Chem. Int. Ed. 2007, 46, 6634–6637.
  2. Berthier, E.; Young, E.W.K.; Beebe, D. Engineers are from PDMS-land, Biologists are from Polystyrenia. Lab Chip 2012, 12, 1224–1237.
  3. Merkel, T.C.; Bondar, V.I.; Nagai, K.; Freeman, B.D.; Pinnau, I. Gas sorption, diffusion, and permeation in Poly(dimethylsiloxane). J. Polym. Sci. Part B Polym. Phys. 2000, 38, 415–434.
  4. Kuddannaya, S.; Bao, J.; Zhang, Y. Enhanced In Vitro biocompatibility of chemically modified Poly(dimethylsiloxane) surfaces for stable adhesion and long-term investigation of brain cerebral cortex cells. ACS Appl. Mater. Interfaces 2015, 7, 25529–25538.
  5. Lee, S.; Shin, H.-J.; Yoon, S.-M.; Yi, D.K.; Choi, J.-Y.; Paik, U. Refractive index engineering of transparent ZrO2–polydimethylsiloxane nanocomposites. J. Mater. Chem. 2008, 18, 1751–1755.
  6. Johnston, I.D.; Tracey, M.C.; Davis, J.B.; Tan, C.K.L. Micro throttle pump employing displacement amplification in an elastomeric substrate. J. Micromech. Microeng. 2005, 15, 1831–1839.
  7. Dardouri, M.; Bettencourt, A.; Martin, V.; Carvalho, F.A.; Santos, C.; Monge, N.; Santos, N.C.; Fernandes, M.H.; Gomes, P.S.; Ribeiro, I.A.C. Using plasma-mediated covalent functionalization of rhamnolipids on polydimethylsiloxane towards the antimicrobial improvement of catheter surfaces. Mater. Sci. Eng. C 2021, 112563.
  8. Kumar, R.; Sahani, A. Role of superhydrophobic coatings in biomedical applications. Mater. Today 2021, 45, 5655–5659.
  9. Wu, X.; Kim, S.-H.; Ji, C.-H.; Allen, M. A solid hydraulically amplified piezoelectric microvalve. J. Micromech. Microeng. 2011, 21, 95003–95011.
  10. Bozukova, D.; Pagnoulle, C.; Jérôme, R.; Jérôme, C. Polymers in modern ophthalmic implants—Historical background and recent advances. Mater. Sci. Eng. R Rep. 2010, 69, 63–83.
  11. Yu, H.; Zhou, G.; Sinha, S.K.; Chau, F.S.; Wang, S. Lens integrated with self-aligned variable aperture using pneumatic actuation method. Sens. Actuators A Phys. 2010, 159, 105–110.
  12. Doutel, E.; Viriato, N.; Carneiro, J.; Campos, J.B.L.M.; Miranda, J.M. Geometrical effects in the hemodynamics of stenotic and non-stenotic left coronary arteries-numerical and in vitro approaches. Int. J. Numer. Methods Biomed. Eng. 2019, 35, e3207.
  13. Usmani, A.; Muralidhar, K. Flow in an intracranial aneurysm model: Effect of parent artery orientation. J. Vis. 2018, 21, 795–818.
  14. Kim, S.-J.; Lee, D.-S.; Kim, I.-G.; Sohn, D.-W.; Park, J.-Y.; Choi, B.-K.; Kim, S.-W. Evaluation of the biocompatibility of a coating material for an implantable bladder volume sensor. Kaohsiung J. Med. Sci. 2012, 28, 123–129.
  15. Carta, R.; Jourand, P.; Hermans, B.; Thoné, J.; Brosteaux, D.; Vervust, T.; Bossuyt, F.; Axisa, F.; Vanfleteren, J.; Puers, R. Design and implementation of advanced systems in a flexible-stretchable technology for biomedical applications. Sens. Actuators A Phys. 2009, 156, 79–87.
  16. Fujii, T. PDMS-based microfluidic devices for biomedical applications. Microelectron. Eng. 2002, 61–62, 907–914.
  17. Raj, M.K.; Chakraborty, S. PDMS microfluidics: A mini review. J. Appl. Polym. Sci. 2020, 137, 48958.
  18. Chen, W.; Lam, R.H.W.; Fu, J. Photolithographic surface micromachining of polydimethylsiloxane (PDMS). Lab Chip 2012, 12, 391–395.
  19. Zhou, J.; Ellis, A.V.; Voelcker, N.H. Recent developments in PDMS surface modification for microfluidic devices. Electrophoresis 2010, 31, 2–16.
  20. Weibel, D.B.; DiLuzio, W.R.; Whitesides, G.M. Microfabrication meets microbiology. Nat. Rev. Microbiol. 2007, 5, 209–218.
  21. Mata, A.; Fleischman, A.J.; Roy, S. Characterization of Polydimethylsiloxane (PDMS) Properties for Biomedical Micro/Nanosystems. Biomed. Microdevices 2005, 7, 281–293.
  22. Ashraf, M.W.; Tayyaba, S.; Afzulpurkar, N. Micro Electromechanical Systems (MEMS) based microfluidic devices for biomedical applications. Int. J. Mol. Sci. 2011, 12, 3648–3704.
  23. Schneider, F.; Fellner, T.; Wilde, J.; Wallrabe, U. Mechanical properties of silicones for MEMS. J. Micromech. Microeng. 2008, 18, 065008.
  24. Bubendorfer, A.; Liu, X.; Ellis, A.V. Microfabrication of PDMS microchannels using SU-8/PMMA moldings and their sealing to polystyrene substrates. Smart Mater. Struct. 2007, 16, 367–371.
  25. Pinto, V.C.; Sousa, P.J.; Cardoso, V.F.; Minas, G. Optimized SU-8 Processing for low-cost microstructures fabrication without cleanroom facilities. Micromachines 2014, 5, 738–755.
  26. Shakeri, A.; Khan, S.; Didar, T.F. Conventional and emerging strategies for the fabrication and functionalization of PDMS-based microfluidic devices. Lab Chip 2021, 21, 3053–3075.
  27. Jo, M.C.; Guldiken, R. Effects of polydimethylsiloxane (PDMS) microchannels on surface acoustic wave-based microfluidic devices. Microelectron. Eng. 2014, 113, 98–104.
  28. Levitt, M.R.; Mandrycky, C.; Abel, A.; Kelly, C.M.; Levy, S.; Chivukula, V.K.; Zheng, Y.; Aliseda, A.; Kim, L.J. Genetic correlates of wall shear stress in a patient-specific 3D-printed cerebral aneurysm model. J. Neurointerv. Surg. 2019, 11, 999–1003.
  29. Doutel, E.; Carneiro, J.; Oliveira, M.; Campos, J.B.L.M.; Miranda, J. Fabrication of 3d mili-scale channels for hemodynamic studies. J. Mech. Med. Biol. 2014, 5, 21.
  30. Doutel, E.; Carneiro, J.; Campos, J.B.L.M.; Miranda, J.M. Experimental and numerical methodology to analyze flows in a coronary bifurcation. Eur. J. Mech. B Fluids 2018, 67, 341–356.
  31. Geoghegan, P.H.; Buchmann, N.A.; Spence, C.J.T.; Moore, S.; Jermy, M. Fabrication of rigid and flexible refractive-index-matched flow phantoms for flow visualisation and optical flow measurements. Exp. Fluids 2012, 52, 1331–1347.
  32. Ford, M.D.; Nikolov, H.N.; Milner, J.S.; Lownie, S.P.; Demont, E.M.; Kalata, W.; Loth, F.; Holdsworth, D.W.; Steinman, D.A. PIV-measured versus CFD-predicted flow dynamics in anatomically realistic cerebral aneurysm models. J. Biomech. Eng. 2008, 130, 21015.
  33. Brindise, M.C.; Rothenberger, S.; Dickerhoff, B.; Schnell, S.; Markl, M.; Saloner, D.; Rayz, V.L.; Vlachos, P.P. Multi-modality cerebral aneurysm haemodynamic analysis: In vivo 4D flow MRI, in vitro volumetric particle velocimetry and in silico computational fluid dynamics. J. R. Soc. Interface 2019, 16, 20190465.
  34. Amili, O.; Golzarian, J.; Coletti, F. In Vitro Study of particle transport in successively bifurcating vessels. Ann. Biomed. Eng. 2019, 47, 2271–2283.
  35. Li, Y.; Verrelli, D.I.; Yang, W.; Qian, Y.; Chong, W. A pilot validation of CFD model results against PIV observations of haemodynamics in intracranial aneurysms treated with flow-diverting stents. J. Biomech. 2020, 100, 109590.
  36. Chivukula, V.K.; Levitt, M.R.; Clark, A.; Barbour, M.C.; Sansom, K.; Johnson, L.; Kelly, C.M.; Geindreau, C.; Rolland du Roscoat, S.; Kim, L.J.; et al. Reconstructing patient-specific cerebral aneurysm vasculature for in vitro investigations and treatment efficacy assessments. J. Clin. Neurosci. Off. J. Neurosurg. Soc. Australas. 2019, 61, 153–159.
  37. Paliwal, N.; Damiano, R.J.; Varble, N.A.; Tutino, V.M.; Dou, Z.; Siddiqui, A.H.; Meng, H. Methodology for Computational Fluid Dynamic Validation for Medical Use: Application to Intracranial Aneurysm. J. Biomech. Eng. 2017, 139, 1210041–12100410.
  38. Rossi de Aguiar, K.M.F.; Nascimento, M.V.; Faccioni, J.L.; Noeske, P.L.M.; Gätjen, L.; Rischka, K.; Rodrigues-Filho, U.P. Urethanes PDMS-based: Functional hybrid coatings for metallic dental implants. Appl. Surf. Sci. 2019, 484, 1128–1140.
  39. Tran, P.A.; Fox, K.; Tran, N. Novel hierarchical tantalum oxide-PDMS hybrid coating for medical implants: One pot synthesis, characterization and modulation of fibroblast proliferation. J. Colloid Interface Sci. 2017, 485, 106–115.
  40. Hijón, N.; Manzano, M.; Salinas, A.; Vallet-Regí, M. Bioactive CaO-SiO2-PDMS coatings on Ti6Al4V substrates. Chem. Mater. 2005, 17, 1591–1596.
  41. Lee, D.S.; Kim, S.J.; Sohn, J.H.; Kim, I.G.; Kim, S.W.; Sohn, D.W.; Kim, J.H.; Choi, B. Biocompatibility of a pdms-coated micro-device: Bladder volume monitoring sensor. Chin. J. Polym. Sci. 2012, 30, 242–249.
  42. Tavakoli, S.; Nemati, S.; Kharaziha, M.; Akbari-Alavijeh, S. Embedding CuO nanoparticles in PDMS-SiO2 coating to improve antibacterial characteristic and corrosion resistance. Colloids Interface Sci. Commun. 2019, 28, 20–28.
  43. Chen, S.; Jones, J.A.; Xu, Y.; Low, H.-Y.; Anderson, J.M.; Leong, K.W. Characterization of topographical effects on macrophage behavior in a foreign body response model. Biomaterials 2010, 31, 3479–3491.
  44. Guo, R.; Liu, J. Implantable liquid metal-based flexible neural microelectrode array and its application in recovering animal locomotion functions. J. Micromech. Microeng. 2017, 27, 104002.
  45. Hassler, C.; Boretius, T.; Stieglitz, T. Polymers for neural implants. J. Polym. Sci. Part B Polym. Phys. 2011, 49, 18–33.
  46. Wolf, M.P.; Salieb-Beugelaar, G.B.; Hunziker, P. PDMS with designer functionalities—Properties, modifications strategies, and applications. Prog. Polym. Sci. 2018, 83, 97–134.
  47. Zhao, J.; Sheadel, D.A.; Xue, W. Surface treatment of polymers for the fabrication of all-polymer MEMS devices. Sens. Actuators A Phys. 2012, 187, 43–49.
  48. Johnston, I.D.; McCluskey, D.K.; Tan, C.K.L.; Tracey, M.C. Mechanical characterization of bulk Sylgard 184 for microfluidics and microengineering. J. Micromech. Microeng. 2014, 24, 035017.
  49. Victor, A.; Ribeiro, J.; Araújo, F.F. Study of PDMS characterization and its applications in biomedicine: A review. J. Mech. Eng. Biomech. 2019, 4, 1–9.
  50. Cardoso, C.; Fernandes, C.S.; Lima, R.; Ribeiro, J. Biomechanical analysis of PDMS channels using different hyperelastic numerical constitutive models. Mech. Res. Commun. 2018, 90, 26–33.
  51. Pan, C.T.; Chen, Y.C.; Lin, P.-H.; Hsieh, C.C.; Hsu, F.T.; Lin, P.-H.; Chang, C.M.; Hsu, J.H.; Huang, J.C. Lens of controllable optical field with thin film metallic glasses for UV-LEDs. Opt. Express 2014, 22, 14411.
  52. Wang, B.; Liu, H.; Zhang, B.; Han, Y.; Shen, C.; Lin, Q.; Chen, H. Development of antibacterial and high light transmittance bulk materials: Incorporation and sustained release of hydrophobic or hydrophilic antibiotics. Colloids Surf. B Biointerfaces 2016, 141, 483–490.
  53. Pan, C.T.; Chen, Y.C.; Chen, Y.J.; Wang, W.C.; Yang, H.C.; Wu, H.C. Compound optical film using gray scale mask embedded with microvoids. Adv. Condens. Matter Phys. 2012, 2012, 942018.
  54. Riehle, N.; Thude, S.; Götz, T.; Kandelbauer, A.; Thanos, S.; Tovar, G.; Lorenz, G. Influence of PDMS molecular weight on transparency and mechanical properties of soft polysiloxane-urea-elastomers for intraocular lens application. Eur. Polym. J. 2018, 101, 190–201.
  55. Sales, F.; Souza, A.; Ariati, R.; Noronha, V.; Giovanetti, E.; Lima, R.; Ribeiro, J. Composite material of PDMS with interchangeable transmittance: Study of optical, mechanical properties and wettability. J. Compos. Sci. 2021, 5, 110.
  56. Mark, J.E. (Ed.) Polymer Data Handbook; Oxford University Press: Oxford, UK, 1999.
  57. The Dow CompanyChemical. SYLGARDTM 184 Silicone Elastomer Technical Datasheet. Silicone Elastomer Technical Data Sheet 2017. Available online: https://consumer.dow.com/en-us/document-viewer.html?ramdomVar=3835418757322904567&docPath=/documents/en-us/productdatasheet/11/11-31/11-3184-sylgard-184-elastomer.pdf (accessed on 20 August 2021).
  58. Hong, J.; Lee, J.; Hong, C.; Shim, S. Effect of dispersion state of carbon nanotube on the thermal conductivity of poly(dimethyl siloxane) composites. Curr. Appl. Phys. 2010, 10, 359–363.
  59. Armani, D.; Liu, C.; Aluru, N. Re-Configurable Fluid Circuits by PDMS Elastomer Micromachining. In Proceedings of the Technical Digest IEEE International MEMS 99 Conference. Twelfth IEEE International Conference on Micro Electro Mechanical Systems (Cat. No.99CH36291), Orlando, FL, USA, 21 January 1999; pp. 222–227.
  60. Müller, A.; Wapler, M.C.; Wallrabe, U. A quick and accurate method to determine the Poisson’s ratio and the coefficient of thermal expansion of PDMS. Soft Matter 2019, 15, 779–784.
  61. Zhang, G.; Sun, Y.; Qian, B.; Gao, H.; Zuo, D. Experimental study on mechanical performance of polydimethylsiloxane (PDMS) at various temperatures. Polym. Test. 2020, 90, 106670.
  62. Gokaltun, A.; Yarmush, M.L.; Asatekin, A.; Usta, O.B. Recent advances in nonbiofouling PDMS surface modification strategies applicable to microfluidic technology. Technology 2017, 5, 1–12.
  63. GRIFFITHS, E. international critical tables of numerical data, physics, chemistry and technology. Nature 1927, 119, 735–738.
  64. Wu, M.H.; Urban, J.P.G.; Cui, Z.; Cui, Z.F. Development of PDMS microbioreactor with well-defined and homogenous culture environment for chondrocyte 3-D culture. Biomed. Microdevices 2006, 8, 331–340.
  65. Tan, S.H.; Nguyen, N.T.; Chua, Y.C.; Kang, T.G. Oxygen plasma treatment for reducing hydrophobicity of a sealed polydimethylsiloxane microchannel. Biomicrofluidics 2010, 4, 032204.
  66. Nakano, H.; Kakinoki, S.; Iwasaki, Y. Long-lasting hydrophilic surface generated on poly(dimethyl siloxane) with photoreactive zwitterionic polymers. Colloids Surf. B Biointerfaces 2021, 205, 111900.
  67. Lee, J.N.; Park, C.; Whitesides, G.M. Solvent compatibility of poly(dimethylsiloxane)-based microfluidic devices. Anal. Chem. 2003, 75, 6544–6554.
  68. Toepke, M.W.; Beebe, D.J. PDMS absorption of small molecules and consequences in microfluidic applications. Lab Chip 2006, 6, 1484–1486.
  69. Bodas, D.; Khan-Malek, C. Hydrophilization and hydrophobic recovery of PDMS by oxygen plasma and chemical treatment-An SEM investigation. Sens. Actuators B Chem. 2007, 123, 368–373.
  70. Makamba, H.; Kim, J.H.; Lim, K.; Park, N.; Hahn, J.H. Surface modification of poly(dimethylsiloxane) microchannels. Electrophoresis 2003, 24, 3607–3619.
  71. Zhou, J.; Khodakov, D.A.; Ellis, A.V.; Voelcker, N.H. Surface modification for PDMS-based microfluidic devices. Electrophoresis 2012, 33, 89–104.
  72. Hemmilä, S.; Cauich-Rodríguez, J.V.; Kreutzer, J.; Kallio, P. Rapid, simple, and cost-effective treatments to achieve long-term hydrophilic PDMS surfaces. Appl. Surf. Sci. 2012, 258, 9864–9875.
  73. Trantidou, T.; Elani, Y.; Parsons, E.; Ces, O. Hydrophilic surface modification of pdms for droplet microfluidics using a simple, quick, and robust method via PVA deposition. Microsyst. Nanoeng. 2017, 3, 16091.
  74. Yang, Y.; Kulangara, K.; Lam, R.T.S.; Dharmawan, R.; Leong, K.W. Effects of Topographical and mechanical property alterations induced by oxygen plasma modification on stem cell behavior. ACS Nano 2012, 6, 8591–8598.
  75. Berdichevsky, Y.; Khandurina, J.; Guttman, A.; Lo, Y.-H. UV/ozone modification of poly(dimethylsiloxane) microfluidic channels. Sens. Actuators B Chem. 2004, 97, 402–408.
  76. Hillborg, H.; Gedde, U.W. Hydrophobicity recovery of polydimethylsiloxane after exposure to corona discharges. Polymer 1998, 39, 1991–1998.
  77. Makamba, H.; Hsieh, Y.Y.; Sung, W.C.; Chen, S.H. Stable permanently hydrophilic protein-resistant thin-film coatings on poly(dimethylsiloxane) substrates by electrostatic self-assembly and chemical cross-linking. Anal. Chem. 2005, 77, 3971–3978.
  78. Boxshall, K.; Wu, M.-H.; Cui, Z.; Cui, Z.; Watts, J.F.; Baker, M.A. Simple surface treatments to modify protein adsorption and cell attachment properties within a poly(dimethylsiloxane) micro-bioreactor. Surf. Interface Anal. 2006, 38, 198–201.
  79. Blättler, T.M.; Pasche, S.; Textor, M.; Griesser, H.J. High salt stability and protein resistance of poly(L-lysine)-g-poly(ethylene glycol) copolymers covalently immobilized via aldehyde plasma polymer interlayers on inorganic and polymeric substrates. Langmuir ACS J. Surf. Colloids 2006, 22, 5760–5769.
  80. Xu, J.; Gleason, K.K. Conformal, amine-functionalized thin films by initiated chemical vapor deposition (iCVD) for hydrolytically stable microfluidic devices. Chem. Mater. 2010, 22, 1732–1738.
  81. Zhang, Z.; Feng, X.; Xu, F.; Liu, X.; Liu, B.F. “Click” chemistry-based surface modification of poly(dimethylsiloxane) for protein separation in a microfluidic chip. Electrophoresis 2010, 31, 3129–3136.
  82. Hu, S.; Ren, X.; Bachman, M.; Sims, C.E.; Li, G.P.; Allbritton, N.L. Surface-directed, graft polymerization within microfluidic channels. Anal. Chem. 2004, 76, 1865–1870.
  83. Ariati, R.; Sales, F.; Souza, A.; Lima, R.A.; Ribeiro, J. Polydimethylsiloxane composites characterization and its applications: A review. Polymers 2021, 13, 4258.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 5.8K
Revisions: 2 times (View History)
Update Date: 12 Jan 2022
1000/1000