Micropump

From Wikipedia, the free encyclopedia
A Ti–Cr–Pt tube (~40 μm long) releases oxygen bubbles when immersed in hydrogen peroxide (catalytic decomposition). Polystyrene spheres (1 μm diameter) were added to study the flow kinetics.[1]
Electrochemical micropump activating the flow of human blood through a 50×100 μm pipe.[2]

Micropumps are devices that can control and manipulate small fluid volumes.[3][4] Although any kind of small pump is often referred to as a micropump, a more accurate definition restricts this term to pumps with functional dimensions in the micrometer range. Such pumps are of special interest in microfluidic research, and have become available for industrial product integration in recent years. Their miniaturized overall size, potential cost and improved dosing accuracy compared to existing miniature pumps fuel the growing interest for this innovative kind of pump.

Note that the below text is very incomplete in terms of providing a good overview of the different micropump types and applications, and therefore please refer to good review articles on the topic.[3][5][6][7]

Introduction and history[edit]

First true micropumps were reported in the mid-1970s,[8] but attracted interest only in the 1980s, when Jan Smits and Harald Van Lintel developed MEMS micropumps.[9] Most of the fundamental MEMS micropump work was done in the 1990s. More recently, efforts have been made to design non-mechanical micropumps that are functional in remote locations due to their non-dependence on external power.

A diagram showing how three microvalves in series can be used to displace fluid. In step (A), fluid is pulled from the inlet into the first valve. Steps (B) – (E) move the fluid to the final valve, before the fluid is expelled towards the outlet in step (F).

Types and technology[edit]

Within the microfluidic world, physical laws change their appearance.[10] As an example, volumetric forces, such as weight or inertia, often become negligible, whereas surface forces can dominate fluidical behaviour,[11] especially when gas inclusion in liquids is present. With only a few exceptions, micropumps rely on micro-actuation principles, which can reasonably be scaled up only to a certain size.

Micropumps can be grouped into mechanical and non-mechanical devices.[12] Mechanical systems contain moving parts, which are usually actuation and microvalve membranes or flaps. The driving force can be generated by utilizing piezoelectric,[13] electrostatic, thermo-pneumatic, pneumatic or magnetic effects. Non-mechanical pumps function with electro-hydrodynamic, electro-osmotic, electrochemical [14] or ultrasonic flow generation, just to name a few of the actuation mechanisms that are currently studied.

Mechanical micropumps[edit]

Diaphragm micropumps[edit]

A diaphragm micropump uses the repeated actuation of a diaphragm to drive a fluid. The membrane is positioned above a main pump valve, which is centered between inlet and outlet microvalves. When the membrane is deflected upwards through some driving force, fluid is pulled into the inlet valve into the main pump valve. The membrane is then lowered, expelling the fluid through the outlet valve. This process is repeated to pump fluid continuously.[6]

Piezoelectric micropumps[edit]

Piezoelectric micropump is one of the most common type of displacement reciprocating diaphragm pumps. Piezoelectric driven micropumps rely on electromechanical property of piezo ceramic to deform in response to applied voltage. Piezoelectric disk attached to the membrane causes diaphragm deflection driven by the external axial electric field thus expanding and contracting the chamber of the micropump.[15] This mechanical strain results in pressure variation in the chamber, which causes inflow and outflow of the fluid. The flow rate is controlled by the polarization limit of the material and the voltage applied on the piezo.[16] In comparison with other actuation principles piezoelectric actuation enables high stroke volume, high actuation force and fast mechanical response, though requiring comparatively high actuation voltage and complex mounting procedure of the piezo ceramic.[9]

The smallest piezoelectric micropump with dimensions of 3.5x3.5x0.6 mm3 was developed by Fraunhofer EMFT[17] the world-renowned research organization with focus on MEMS and Microsystem technologies. The micropump consists of three silicon layers, one of which as a pump diaphragm confines the pump chamber from above, while two others represent middle valve chip and bottom valve chip. Openings of the passive flap valves at the inlet and outlet are oriented according to the flow direction. The pump diaphragm expands with application of a negative voltage to the piezo thus creating negative pressure to suck the fluid into the pump chamber. While positive voltage vice versa drives the diaphragm down, which results in overpressure opening outlet valve and forcing the fluid out of the chamber.

Back pressure performance of 3.5x3.5mm2 silicon piezoelectric driven micropump
Openings of the passive flap valves at the inlet and outlet are oriented according to the flow direction. The pump diaphragm expands with application of a negative voltage to the piezo thus creating negative pressure to suck the fluid into the pump chamber in supply mode. While positive voltage drives the diaphragm down, which results in opening outlet valve due to overpressure in pump mode.


Currently mechanical micropump technology extensively uses Silicon and Glass based micromachining processes for fabrication. Among the common microfabrication processes, the following techniques can be named: photolithography, anisotropic etching, surface micromachining and bulk micromachining of silicon.[16] Silicon micromachining has numerous advantages that facilitate the technology widespread in high performance applications as, for example, in drug delivery.[9] Thus, silicon micromachining allows high geometric precision and long-term stability, since mechanically moving parts, e.g. valve flaps, do not exhibit wear and fatigue. As an alternative to silicon polymer-based materials like PDMS, PMMA, PLLA, etc. can be used due to the superior strength, enhanced structural properties, stability and inexpensiveness. Silicon micropumps at Fraunhofer EMFT are manufactured by silicon micromachining technology.[18] Three monocrystalline silicon wafers (100 oriented) are structured by doublesided lithography and etched by silicon wet etching (using potassium hydroxide solution KOH). The connection between the structured wafer layers is realized by silicon fusion bond. This bonding technology needs very smooth surfaces (roughness lower than 0.3 nm) and very high temperatures (up to 1100 °C) to perform a direct silicon–silicon bond between the wafer layers. Absence of the bonding layer allows definition of the vertical pump design parameters. Additionally, the bonding layer might be affected by the pumped medium.

The compression ratio of the micropump as one of the critical performance indicator is defined as the ratio between the stroke volume, i.e. fluid volume displaced by the pump membrane over the course of the pump cycle, and the dead volume, i.e. the minimum fluid volume remaining in the pump chamber in pumping mode.[15]

The compression ratio defines the bubble tolerance and the counter pressure capability of the micropumps. Gas bubbles within chamber hinder micropump operation as due to the damping properties of the gas bubbles the pressure peaks (∆P) in the pump chamber decreases, while due to the surface properties the critical pressure (∆Pcrit) that opens passive valves increases.[19] The compression ratio of Fraunhofer EMFT micropumps reaches the value of 1, which implies self-priming capability and bubble tolerance even at challenging outlet pressure conditions. Large compression ratio is achieved thanks to special patented technique of piezo mounting, when electrical voltage is applied on the electrodes on the top and bottom of the piezoelectric ceramic during the curing process of the adhesive used for the piezo mounting. Considerable reduction of the dead volume resulted from predeflected actuators along with shallow fabricated pump chamber heights increases the compression ratio.

Peristaltic micropumps[edit]

A peristaltic micropump is a micropump composed of at least three microvalves in series. These three valves are opened and closed sequentially in order to pull fluid from the inlet to the outlet in a process known as peristalsis.[20]

Non-mechanical micropumps[edit]

Valveless micropumps[edit]

Static valves are defined as valves which have fixed geometry without any moving parts. These valves provide flow rectification through addition of energy (active) or inducing desired flow behavior by fluid inertia (passive). Two most common types of static geometry passive valves are Diffuser-Nozzle Elements [21][22] and Tesla valves. Micropumps having nozzle-diffuser elements as flow rectification device are commonly known as Valveless Micropumps.

Capillary pumps[edit]

In microfluidics, capillary pumping plays an important role because the pumping action does not require external actuation power. Glass capillaries and porous media, including nitrocellulose paper and synthetic paper,[23] can be integrated into microfluidic chips. Capillary pumping is widely used in lateral flow testing. Recently, novel capillary pumps, with a constant pumping flow rate independent of the liquid viscosity and surface energy,[24][25][26][27] were developed, which have a significant advantage over the traditional capillary pump (of which the flow behaviour is Washburn behaviour, namely the flow rate is not constant) because their performance does not depend on the sample viscosity.

Chemically powered pumps[edit]

Chemically powered non-mechanical pumps have been fabricated by affixing nanomotors to surfaces, driving fluid flow through chemical reactions. A wide variety of pumping systems exist including biological enzyme based pumps,[28][29][30][31][32][33] organic photocatalyst pumps,[34] and metal catalyst pumps.[31][35] These pumps generate flow through a number of different mechanisms including self-diffusiophoresis, electrophoresis, bubble propulsion and the generation of density gradients.[29][32][36] Moreover, these chemically powered micropumps can be used as sensors for the detection of toxic agents.[30][37]

Light-powered pumps[edit]

Another class of non-mechanical pumping is light-powered pumping.[38][39] Certain nanoparticles are able to convert light from a UV source to heat which generates convective pumping. These kinds of pumps are possible with titanium dioxide nanoparticles and the speed of pumping can be controlled by both the intensity of the light source and the concentration of particles.[40]

Applications[edit]

Micropumps have potential industrial applications, such as delivery of small amounts of glue during manufacturing processes, and biomedical applications, including portable or implanted drug delivery devices. Bio-inspired applications include a flexible electromagnetic micropump using magnetorheological elastomer to replace lymphatic vessels.[41] Chemically powered micropumps also demonstrate potential for applications in chemical sensing in terms of detecting chemical warfare agents and environmental hazards, such as mercury and cyanide.[30]

Considering contemporary state of air pollution one of the most promising applications for micropump lies in enhancement of gas and particulate matter sensors for monitoring personal air quality. Thanks to MEMS fabrication technology, gas sensors based on MOS, NDIR, electrochemical principles could be miniaturized to fit portable devices as well as smartphones and wearables. Application of the Fraunhofer EMFT piezoelectric micropump reduces reaction time of the sensor up to 2 seconds through fast sampling of the ambient air.[42] This is explained by the fast convection that takes place when micropump drives the air towards the sensor, while in absence of the micropump due to slow diffusion sensor response is delayed for several minutes. The current alternative to the micropump – the fan – has numerous drawbacks. Unable to achieve substantial negative pressure fan cannot overcome pressure drop at the filter diaphragm. Further, the gas molecules and particles can easily re-adhere to the sensor surface and its housing, which in time results in sensor drift.

Additionally inbuilt micropump facilitates sensor regeneration and thus resolves saturation issues by expelling gas molecules out of the sensor surface. Breath analysis is related field of use for the gas sensor that is empowered by micropump. Micropump can advance remote diagnostic and monitoring of gastrointestinal tract and pulmonary diseases, diabetes, cancer etc. by means of portable devices within telemedicine programs.

The promising application for MEMS micropumps lies in drug delivery systems for diabetes-, tumor-, hormone-, pain and ocular therapy in forms of ultra-thin patches, targeted delivery within implantable systems or intelligent pills. Piezoelectric MEMS micropumps can replace traditional peristaltic or syringe pumps for intravenous, subcutaneous, arterial, ocular drug injection. Drug delivery application does not require high flow rates, however, the micropumps are supposed to be precise in delivering small doses and demonstrate back pressure independent flow.[16] Due to biocompatibility and miniature size, silicon piezoelectric micropump can be implanted on the eyeball to treat glaucoma or phthisis. Since under these conditions the eye loses its ability to ensure outflow or production of aqueous humour, the implanted micropump developed by Fraunhofer EMFT with the flow rate of 30 µl/s facilitates proper flow of the fluid without restricting or creating any inconvenience to the patient.[43] Another health issue to be solved by micropump is bladder incontinence. Artificial sphincter technology based on the titanium micropump ensures continence by automatically adjusting the pressure during laughter or coughing. The urethra is opened and closed by means of a fluid-filled sleeve that is regulated by the micropump.[44]

Micropump can facilitate scent scenario for consumer, medical, defense, first responder applications etc. to enhance the effect of with ubiquitous picture scenarios (movies) and sound scenarios (music). Microdosing device with several scent reservoirs that are mounted near the nose can release 15 different scent impressions in 1 min.[18] Advantage of the micropump lies in the possibility to smell sequence of scents without different odors being mixed. The system ensures an appropriate dose of the scent to be detected by the user only as soon as scent molecules are delivered. Numerous applications are possible with micropump for scent-dosing: tasters training (wine, food), learning programs, psychotherapy, anosmia treatment, first responder training etc. to facilitate full immersion in the desired environment.

Within analytical systems, the micropump can be for lab-on-chip applications, HPLC and Gas Chromatography systems etc. For the latter micropumps are required to ensure accurate delivery and flow of gases. Since the compressibility of the gases is challenging, the micropump must possess high compression ratio.[16]

Among other applications, the following fields can be named: dosage systems for small quantity of lubricants, fuel dosing systems, micro pneumatics, micro hydraulic systems and dosage systems in production processes, liquid handling (cushion pipettes, microliter plates).[45]

See also[edit]

References[edit]

  1. ^ Solovev, Alexander A.; Sanchez, Samuel; Mei, Yongfeng; Schmidt, Oliver G. (2011). "Tunable catalytic tubular micro-pumps operating at low concentrations of hydrogen peroxide". Physical Chemistry Chemical Physics. 13 (21): 10131–5. Bibcode:2011PCCP...1310131S. doi:10.1039/C1CP20542K. PMID 21505711. S2CID 21754449.
  2. ^ Chiu, S. H.; Liu, C. H. (2009). "An air-bubble-actuated micropump for on-chip blood transportation". Lab on a Chip. 9 (11): 1524–33. doi:10.1039/B900139E. PMID 19458858. S2CID 38015356.
  3. ^ a b Bußmann, Agnes Beate; Grünerbel, Lorenz Maximilian; Durasiewicz, Claudia Patricia; Thalhofer, Thomas Alexander; Wille, Axel; Richter, Martin (2021-10-15). "Microdosing for drug delivery application—A review". Sensors and Actuators A: Physical. 330: 112820. doi:10.1016/j.sna.2021.112820. ISSN 0924-4247.
  4. ^ Laser, D. J.; Santiago, J. G. (2004). "A review of micropumps". Journal of Micromechanics and Microengineering. 14 (6): R35. Bibcode:2004JMiMi..14R..35L. doi:10.1088/0960-1317/14/6/R01. ISSN 0960-1317. S2CID 35703576.
  5. ^ Nguyen; et al. (2002). "MEMS-Micropumps: A Review". Journal of Fluids Engineering. 124 (2): 384–392. doi:10.1115/1.1459075.
  6. ^ a b Iverson; et al. (2008). "Recent advances in microscale pumping technologies: a review and evaluation". Microfluid Nanofluid. 5 (2): 145–174. doi:10.1007/s10404-008-0266-8. S2CID 44242994.
  7. ^ Amirouche; et al. (2009). "Current micropump technologies and their biomedical applications". Microsystem Technologies. 15 (5): 647–666. doi:10.1007/s00542-009-0804-7. S2CID 108575489.
  8. ^ Thomas, L. J. and Bessman, S. P. (1975) "Micropump powered by piezoelectric disk benders", U.S. patent 3,963,380
  9. ^ a b c Woias, P (2005). "Micropumps – past progress and future prospects". Sensors and Actuators B. 105 (1): 28–38. doi:10.1016/j.snb.2004.02.033.
  10. ^ Order from Chaos Archived 2008-07-23 at the Wayback Machine, The CAFE Foundation
  11. ^ Thomas, D. J.; Tehrani, Z.; Redfearn, B. (2016-01-01). "3-D printed composite microfluidic pump for wearable biomedical applications". Additive Manufacturing. 9: 30–38. doi:10.1016/j.addma.2015.12.004. ISSN 2214-8604.
  12. ^ Wang, Yao-Nan; Fu, Lung-Ming (5 August 2018). "Micropumps and biomedical applications – A review". Microelectronic Engineering. 195: 121–138. doi:10.1016/j.mee.2018.04.008. S2CID 139917725.
  13. ^ Farshchi Yazdi, Seyed Amir Fouad; Corigliano, Alberto; Ardito, Raffaele (2019-04-18). "3-D Design and Simulation of a Piezoelectric Micropump". Micromachines. 10 (4): 259. doi:10.3390/mi10040259. ISSN 2072-666X. PMC 6523882. PMID 31003481.
  14. ^ Neagu, C.R.; Gardeniers, J.G.E.; Elwenspoek, M.; Kelly, J.J. (1996). "An electrochemical microactuator: principle and first results". Journal of Microelectromechanical Systems. 5 (1): 2–9. doi:10.1109/84.485209.
  15. ^ a b Laser and Santiago (2004). "A review of micropumps". J. Micromech. Microeng. 14 (6): R35–R64. Bibcode:2004JMiMi..14R..35L. doi:10.1088/0960-1317/14/6/R01. S2CID 35703576.
  16. ^ a b c d Mohith, S.; Karanth, P. Navin; Kulkarni, S. M. (2019-06-01). "Recent trends in mechanical micropumps and their applications: A review". Mechatronics. 60: 34–55. doi:10.1016/j.mechatronics.2019.04.009. ISSN 0957-4158. S2CID 164712738.
  17. ^ "Miniaturized micro patch pump – Fraunhofer EMFT". Fraunhofer Research Institution for Microsystems and Solid State Technologies EMFT. 6 November 2019. Retrieved 2019-12-03.
  18. ^ a b Richter, Martin (2017). "Microdosing of Scent". In Buettner, Andrea (ed.). Handbook of Odor. Springer International Publishing. pp. 1081–1097. ISBN 978-3-319-26930-6.
  19. ^ Richter, M.; Linnemann, R.; Woias, P. (1998-06-15). "Robust design of gas and liquid micropumps". Sensors and Actuators A: Physical. Eurosensors XI. 68 (1): 480–486. doi:10.1016/S0924-4247(98)00053-3. ISSN 0924-4247.
  20. ^ Smits, Jan G. (1990). "Piezoelectric micropump with three valves working peristaltically". Sensors and Actuators A: Physical. 21 (1–3): 203–206. doi:10.1016/0924-4247(90)85039-7.
  21. ^ Stemme and Stemme (1993). "A valveless diffuser/nozzle-based fluid pump". Sensors and Actuators A: Physical. 39 (2): 159–167. doi:10.1016/0924-4247(93)80213-Z.
  22. ^ van der Wijngaart (2001). "A valve-less diffuser micropump for microfluidic analytical systems". Sensors and Actuators B: Chemical. 72 (3): 259–265. doi:10.1016/S0925-4005(00)00644-4.
  23. ^ Jonas Hansson; Hiroki Yasuga; Tommy Haraldsson; Wouter van der Wijngaart (2016). "Synthetic microfluidic paper: high surface area and high porosity polymer micropillar arrays". Lab on a Chip. 16 (2): 298–304. doi:10.1039/C5LC01318F. PMID 26646057.
  24. ^ Weijin Guo; Jonas Hansson; Wouter van der Wijngaart (2016). "Viscosity Independent Paper Microfluidic Imbibition" (PDF). MicroTAS 2016, Dublin, Ireland.
  25. ^ Weijin Guo; Jonas Hansson; Wouter van der Wijngaart (2016). "Capillary Pumping Independent of Liquid Sample Viscosity". Langmuir. 32 (48): 12650–12655. doi:10.1021/acs.langmuir.6b03488. PMID 27798835. S2CID 24662688.
  26. ^ Weijin Guo; Jonas Hansson; Wouter van der Wijngaart (2017). Capillary pumping with a constant flow rate independent of the liquid sample viscosity and surface energy. 2017 IEEE 30th International Conference on Micro Electro Mechanical Systems (MEMS). IEEE MEMS 2017, Las Vegas, US. pp. 339–341. doi:10.1109/MEMSYS.2017.7863410. ISBN 978-1-5090-5078-9. S2CID 13219735.{{cite conference}}: CS1 maint: location (link)
  27. ^ Weijin Guo; Jonas Hansson; Wouter van der Wijngaart (2018). "Capillary pumping independent of the liquid surface energy and viscosity". Microsystems & Nanoengineering. 4 (1): 2. Bibcode:2018MicNa...4....2G. doi:10.1038/s41378-018-0002-9. PMC 6220164. PMID 31057892.
  28. ^ Sengupta, S.; Patra, D.; Ortiz-Rivera, I.; Agrawal, A.; Shklyaev, S.; Dey, K. K.; Córdova-Figueroa, U.; Mallouk, T. E.; Sen, A. (2014). "Self-powered enzyme micropumps". Nature Chemistry. 6 (5): 415–422. Bibcode:2014NatCh...6..415S. doi:10.1038/nchem.1895. PMID 24755593. S2CID 14639241.
  29. ^ a b Ortiz-Rivera, I.; Shum, H.; Agrawal, A.; Balazs, A. C.; Sen, A. (2016). "Convective flow reversal in self-powered enzyme micropumps". Proceedings of the National Academy of Sciences. 113 (10): 2585–2590. Bibcode:2016PNAS..113.2585O. doi:10.1073/pnas.1517908113. PMC 4791027. PMID 26903618.
  30. ^ a b c Ortiz-Rivera, I.; Courtney, T.; Sen, A. (2016). "Enzyme Micropump-Based Inhibitor Assays". Advanced Functional Materials. 26 (13): 2135–2142. doi:10.1002/adfm.201504619. S2CID 101206241.
  31. ^ a b Das, S.; Shklyaev, O. E.; Altemose, A.; Shum, H.; Ortiz-Rivera, I.; Valdez, L.; Mallouk, T. E.; Balazs, A. C.; Sen, A. (2017-02-17). "Harnessing catalytic pumps for directional delivery of microparticles in microchambers". Nature Communications. 8: 14384. Bibcode:2017NatCo...814384D. doi:10.1038/ncomms14384. ISSN 2041-1723. PMC 5321755. PMID 28211454.
  32. ^ a b Valdez, L.; Shum, H.; Ortiz-Rivera, I.; Balazs, A. C.; Sen, A. (2017). "Solutal and thermal buoyancy effects in self-powered phosphatase micropumps". Soft Matter. 13 (15): 2800–2807. Bibcode:2017SMat...13.2800V. doi:10.1039/C7SM00022G. PMID 28345091. S2CID 22257211.
  33. ^ Maiti, Subhabrata; Shklyaev, Oleg E.; Balazs, Anna C.; Sen, Ayusman (2019-03-12). "Self-Organization of Fluids in a Multienzymatic Pump System". Langmuir. 35 (10): 3724–3732. doi:10.1021/acs.langmuir.8b03607. ISSN 0743-7463. PMID 30721619. S2CID 73415792.
  34. ^ Yadav, V.; Zhang, H.; Pavlick, R.; Sen, A. (2012). "Triggered "On/Off" Micropumps and Colloidal Photodiode". Journal of the American Chemical Society. 134 (38): 15688–15691. doi:10.1021/ja307270d. PMID 22971044.
  35. ^ Solovev, A. A.; Sanchez, S.; Mei, Y.; Schmidt, O. G. (2011). "Tunable catalytic tubular micro-pumps operating at low concentrations of hydrogen peroxide". Physical Chemistry Chemical Physics. 13 (21): 10131–10135. Bibcode:2011PCCP...1310131S. doi:10.1039/c1cp20542k. PMID 21505711. S2CID 21754449.
  36. ^ Yadav, V.; Duan, W.; Butler, P. J.; Sen, A. (2015). "Anatomy of Nanoscale Propulsion". Annual Review of Biophysics. 44 (1): 77–100. doi:10.1146/annurev-biophys-060414-034216. PMID 26098511.
  37. ^ Zhao, Xi; Gentile, Kayla; Mohajerani, Farzad; Sen, Ayusman (2018-10-16). "Powering Motion with Enzymes". Accounts of Chemical Research. 51 (10): 2373–2381. doi:10.1021/acs.accounts.8b00286. ISSN 0001-4842. PMID 30256612. S2CID 52845451.
  38. ^ Li, Mingtong; Su, Yajun; Zhang, Hui; Dong, Bin (2018-04-01). "Light-powered direction-controlled micropump". Nano Research. 11 (4): 1810–1821. doi:10.1007/s12274-017-1799-5. ISSN 1998-0000. S2CID 139110468.
  39. ^ Yue, Shuai; Lin, Feng; Zhang, Qiuhui; Epie, Njumbe; Dong, Suchuan; Shan, Xiaonan; Liu, Dong; Chu, Wei-Kan; Wang, Zhiming; Bao, Jiming (2019-04-02). "Gold-implanted plasmonic quartz plate as a launch pad for laser-driven photoacoustic microfluidic pumps". Proceedings of the National Academy of Sciences. 116 (14): 6580–6585. Bibcode:2019PNAS..116.6580Y. doi:10.1073/pnas.1818911116. ISSN 0027-8424. PMC 6452654. PMID 30872482.
  40. ^ Tansi, Benjamin M.; Peris, Matthew L.; Shklyaev, Oleg E.; Balazs, Anna C.; Sen, Ayusman (2019). "Organization of Particle Islands through Light-Powered Fluid Pumping". Angewandte Chemie International Edition. 58 (8): 2295–2299. doi:10.1002/anie.201811568. ISSN 1521-3773. PMID 30548990. S2CID 56484282.
  41. ^ Behrooz, M. & Gordaninejad, F. (2014). "A flexible magnetically-controllable fluid transport system". In Liao, Wei-Hsin (ed.). Active and Passive Smart Structures and Integrated Systems 2014. Vol. 9057. pp. 90572Q. doi:10.1117/12.2046359. S2CID 17879262.
  42. ^ "Warnung vor zu viel Feinstaub per Handy". AZ-Online (in German). 2 August 2017. Retrieved 2019-12-04.
  43. ^ "Miniaturpumpe regelt Augeninnendruck". www.labo.de (in German). 3 July 2015. Retrieved 2020-01-13.
  44. ^ "Artificial sphincter system with microfluid actuators – Fraunhofer EMFT". Fraunhofer Research Institution for Microsystems and Solid State Technologies EMFT. Retrieved 2020-01-13.
  45. ^ "Micro Dosing – Fraunhofer EMFT". Fraunhofer Research Institution for Microsystems and Solid State Technologies EMFT. Retrieved 2020-01-13.