Multielectrode array

Jump to navigation Jump to search


WikiDoc Resources for Multielectrode array

Articles

Most recent articles on Multielectrode array

Most cited articles on Multielectrode array

Review articles on Multielectrode array

Articles on Multielectrode array in N Eng J Med, Lancet, BMJ

Media

Powerpoint slides on Multielectrode array

Images of Multielectrode array

Photos of Multielectrode array

Podcasts & MP3s on Multielectrode array

Videos on Multielectrode array

Evidence Based Medicine

Cochrane Collaboration on Multielectrode array

Bandolier on Multielectrode array

TRIP on Multielectrode array

Clinical Trials

Ongoing Trials on Multielectrode array at Clinical Trials.gov

Trial results on Multielectrode array

Clinical Trials on Multielectrode array at Google

Guidelines / Policies / Govt

US National Guidelines Clearinghouse on Multielectrode array

NICE Guidance on Multielectrode array

NHS PRODIGY Guidance

FDA on Multielectrode array

CDC on Multielectrode array

Books

Books on Multielectrode array

News

Multielectrode array in the news

Be alerted to news on Multielectrode array

News trends on Multielectrode array

Commentary

Blogs on Multielectrode array

Definitions

Definitions of Multielectrode array

Patient Resources / Community

Patient resources on Multielectrode array

Discussion groups on Multielectrode array

Patient Handouts on Multielectrode array

Directions to Hospitals Treating Multielectrode array

Risk calculators and risk factors for Multielectrode array

Healthcare Provider Resources

Symptoms of Multielectrode array

Causes & Risk Factors for Multielectrode array

Diagnostic studies for Multielectrode array

Treatment of Multielectrode array

Continuing Medical Education (CME)

CME Programs on Multielectrode array

International

Multielectrode array en Espanol

Multielectrode array en Francais

Business

Multielectrode array in the Marketplace

Patents on Multielectrode array

Experimental / Informatics

List of terms related to Multielectrode array


Overview

Multielectrode arrays (MEAs) or microelectrode arrays are devices that contain multiple plates or shanks through which neural signals are obtained or delivered, essentially serving as neural interfaces that connect neurons to electronic circuitry. There are two general classes of MEAs: implantable MEAs, used in vivo, and non-implantable MEAs, used in vitro.

Theory

Neurons and muscle cells create ion currents through their membranes when excited, causing a change in voltage both inside and outside the cell. When recording, the electrodes on an MEA transduce the change in voltage from the environment carried by ions into currents carried by electrons (electronic currents). When stimulating, electrodes transduce electronic currents into ionic currents through the media. This triggers the voltage-gated ion channels on the membranes of the excitable cells, causing the cell to depolarize and trigger an action potential if it a neuron or a twitch if it is a muscle cell[citation needed].

The size and shape of a recorded signal depend upon several factors: the nature of the medium in which the cell or cells are located (e.g. the medium’s conductivity, capacitance, and homogeneity); the nature of contact between the cells and the MEA electrode (e.g. area of contact and tightness); the nature of the MEA electrode itself (e.g. its geometry, impedance, and noise); the analog signal processing (e.g. the system’s gain, bandwidth, and behavior outside of cutoff frequencies); and the data sampling properties (e.g. sampling rate and digital signal processing)[1]. For the recording of a single cell that partially covers a planar electrode, the voltage at the contact pad is approximately equal to the voltage of the overlapping region of the cell and electrode multiplied by the ratio the surface area of the overlapping region to the area of the entire electrode, or:

<math>V_{pad}=V_{overlap}\times\frac{A_{overlap}}{A_{electrode}}</math>

assuming the area around an electrode is well-insulated and has a very small capacitance associated with it[1]. The equation above, however, relies on modeling the electrode, cells, and their surroundings as an equivalent circuit diagram. An alternative means of predicting cell-electrode behavior is by modeling the system using a geometry-based finite element analysis in an attempt to circumvent the limitations of oversimplifying the system in a lumped circuit element diagram[2].

An MEA can be used to perform electrophysiological experiments on tissue slices or dissociated cell cultures. With acute tissue slices, the connections between the cells within the tissue slices prior to extraction and plating are more or less preserved, while the intercellular connections in dissociated cultures are destroyed prior to plating. With dissociated neuronal cultures, the neurons spontaneously form networks[3].

It can be seen that the voltage amplitude an electrode experiences is inversely related to the distance from which a cell depolarizes[4]. Thus, it may be necessary for the cells to be cultured or otherwise placed as close to the electrodes as possible. With tissue slices, a layer of electrically passive dead cells form around the site of incision due to edema[5]. A way to deal with this is by fabricating an MEA with three-dimensional electrodes fabricated by masking and chemical etching. These 3-D electrodes penetrate the dead cell layer of the slice tissue, decreasing the distance between live cells and the electrodes[6]. In dissociated cultures, proper adherence of the cells to the MEA substrate is important for getting robust signals.

History

The first implantable arrays were microwire arrays developed in the 1950s[7]. The first experiment involving the use of an array of planar electrodes to record from cultured cells was conducted in 1972 by C.A. Thomas, Jr. and his colleagues[4]. The experimental setup used a 2 x 15 array of gold electrodes plated with platinum black, each spaced 100 µm apart from each other. Myocytes harvested from embryonic chicks were dissociated and cultured onto the MEAs, and signals up to 1 mV high in amplitude were recorded[8]. MEAs were constructed and used to explore the electrophysiology of snail ganglia independently by G. Gross and his colleages in 1977 without prior knowledge of Thomas and his colleagues’ work[4]. In 1982, Gross observed spontaneous electrophysiological activity from dissociated spinal cord neurons, and found that activity was very dependent on temperature. Below about 30˚C signal amplitudes decrease rapidly to relatively small value at room temperature[4].

Before the 1990’s, significant entry barriers existed for new laboratories that sought to conduct MEA research due to the custom MEA fabrication and software they had to develop[3]. However, with the advent of affordable computing power[1] and commercial MEA hardware and software[3], many other laboratories were able to undertake research using MEAs.

Types

Microelectrode arrays can be divided up into subcategories based on their potential use: in vitro and in vivo arrays.

Types of in vitro arrays

File:MEAinHand.jpg
An in vitro MEA

The standard type of in vitro MEA comes in a pattern of 8 x 8 or 6 x 10 electrodes. Electrodes are typically composed of indium tin oxide or titanium and have diameters between 10 and 30 μm. These arrays are normally used for single-cell cultures or acute brain slices [1].

One challenge among in vitro MEAs has been imaging them with microscopes that use high power lenses, requiring low working distances on the order of micrometers. In order to avoid this problem, “Thin”-MEAs have been created using cover slip glass. These arrays are approximately 180 μm allowing them to be used with high-power lenses [1] [9].

In another special design, 60 electrodes are split into 6 x 5 arrays separated by 500 μm. Electrodes within a group are separated by 30 um with diameters of 10 μm. Arrays such as this are used to examine local responses of neurons while also studying functional connectivity of organotypic slices [1] [10].

Spatial resolution is one of the key advantages of MEAs and allows signals sent over a long distance to be taken with higher precision when a high-density MEA is used. These arrays usually have a square grid pattern of 256 electrodes that cover an area of 2.8 by 2.8 mm [1].

In order to obtain quality signals electrodes and tissue must be in close contact with one another. The perforated MEA design applies negative pressure to openings in the substrate so that tissue slices can be positioned on the electrodes to enhance contact and recorded signals [1].

The MEA60 System by MULTI CHANNEL SYSTEMS (MCS) is used for a variety of application. It includes a 60-channel amplifier reduces noise level in the retrieved signal so that lower range voltages can be recorded [1] [11].

Types of in vivo arrays

The three major categories of implantable MEAs are microwire, silicon- based, and flexible microelectrode arrays. Microwire MEAs are largely made of stainless steel or tungsten and they can be used to estimate the position of individual recorded neurons by triangulation. Silicon-based microelectrode arrays include two specific models: the Michigan and Utah arrays. Michigan arrays allow a higher density of sensors for implantation as well as a higher spatial resolution than microwire MEAs. They also allow signals to be obtained along the length of the shank, rather than just at the ends of the shanks. In contrast to Michigan arrays, Utah arrays are 3-D, consisting of 100 conductive silicon needles. However, in a Utah array signals are only received from the tips of each electrode, which limits the amount of information that can be obtained at one time. Furthermore, Utah arrays are manufactured with set dimensions and parameters while the Michigan array allows for more design freedom. Flexible arrays, made with polyimide, parylene, or benzocyclobutene, provide an advantage over rigid microelectrode arrays because they provide a closer mechanical match, as the Young’s modulus of silicon is much larger than that of brain tissue, contributing to shear-induced inflammation [7].

Data Processing Methods

The fundamental unit of communication of neurons is, electrically, at least, the action potential. This all-or-nothing phenomenon is believed to originate at the axon hillock,[12] resulting in a depolarization of the intracellular environment which propagates down the axon. This ion flux through the cellular membrane generates a sharp change in voltage in the exracellular environment, which is what the MEA electrodes ultimately detect. Thus, voltage spike counting and sorting is often used in research to characterize network activity.

Advantages

In general, the major strengths of in vitro arrays when compared to more traditional methods such as patch clamping include [13]:

  • Allowing the placement of multiple electrodes at once rather than individually
  • The ability to set up controls within the same experimental setup (by using one electrode as a control and others as experimental)
  • The ability to select different recordings sites within the array
  • The ability to simultaneously receive data from multiple sites

Furthermore, in vitro arrays are non-invasive when compared to patch clamping because they do not require breeching of the cell membrane.

With respect to in vivo arrays however, the major advantage over patch clamping is the high special resolution. Implantable arrays allow signals to be obtained from individual neurons enabling information such as position or velocity of motor movement that can be used to control a prosthetic device.

Disadvantages

In vitro MEAs are less suited for recording and stimulating single cells due to their low spatial resolution compared to patch clamp and dynamic clamp systems. The complexity of signals an MEA electrode could effectively transmit to other cells is limited compared to the capabilities of dynamic clamps.

There are also several biological responses to implantation of a microelectrode array, particularly in regards to chronic implantation. Most notable among these effects are neuronal cell loss, glial scarring, and a drop in the number of functioning electrodes [14]. The tissue response to implantation is dependent among many factors including size of the MEA shanks, distance between the shanks, MEA material composition, and time period of insertion. The tissue response is typically divided into short term and long term response. The short term response occurs within hours of implantation and begins with an increased population of astrocytes and glial cells surrounding the device. The recruited microglia then initiate inflammation and a process of phagocytosis of the foreign material begins. Over time, the astrocytes and microglia recruited to the device begin to accumulate, forming a sheath surrounding the array that extends tens of micrometres around the device. This not only increases the space between electrode probes, but also insulates the electrodes and increases impedance measurements. Problems with chronic implantation of arrays have been a driving force in the research of these devices. One novel study examined the neurodegenerative effects of inflammation caused by chronic implantation [15]. Immunohistochemical markers showed a surprising presence of hyperphosphorylated tau, an indicator of Alzheimer’s disease, near the electrode recording site. The phagocytosis of electrode material also brings to question the issue of a biocompatibility response, which research suggests has been minor and becomes almost nonexistent after 12 weeks in vivo. Research to minimize the negative effects of device insertion include surface coating of the devices with polymers such as laminin as well as coating with drug eluting substances [16].

Applications

In vitro Applications

File:Animat Drawing.jpg
Schematic for a neurally controlled animat

The nature of dissociated neuronal networks does not seem to change or diminish the character of its pharmacological response when compared to in vivo models, suggesting that MEAs can be used to study pharmacological effects on dissociated neuronal cultures in a more simple, controlled environment[17]. A number of pharmacological studies using MEAs on dissociated neuronal networks, e.g. studies with ethanol[18].

MEAs have been used to interface neuronal networks with non-biological systems as a controller. For example, a neural-computer interface can be created using MEAs. Dissociated rat cortical neurons were integrated into a closed stimulus-response feedback loop to control an animat in a virtual environment[19].

A closed-loop stimulus-response system has also been constructed using an MEA by Dr. Potter, Dr. Mandhavan, and Dr. DeMarse[20], and by Mark Hammond, Kevin Warwick, and Ben Whalley in the University of Reading. About 300,000 dissociated rat neurons were plated on an MEA, which was connected to motors and ultrasound sensors on a robot, and was conditioned to avoid obstacles when sensed[21].

MEAs have been used to observe network firing in hippocampal slices[22].

In vivo Applications

There are several implantable interfaces that are currently available for consumer use including deep brain stimulators, cochlear implants, and cardiac pacemakers. Deep brain stimulation (DBS) has been effective at treating movement disorders such as Parkinson’s disease [23], and cochlear implants have helped many to improve their hearing by assisting stimulation of the auditory nerve. Because of their remarkable potential, MEAs are a prominent area of neuroscience research. Research suggests that MEAs may provide insight into processes such as memory formation and perception and may also hold therapeutic value for conditions such as epilepsy, depression, and obsessive-compulsive disorder. Clinical trials using interface devices for restoring motor control after spinal cord injury or as treatment for ALS have been initiated by CyberKinetics Neurotechnology Systems in a project entitled BrainGate (see video demo: BrainGate). MEAs provide the high resolution necessary to record time varying signals, giving them the ability to be used to both control and obtain feedback from prosthetic devices, as was shown by Kevin Warwick, Mark Gasson and Peter Kyberd[24],[25]. Research suggests that MEA use may be able to assist in the restoration of vision by stimulating the optic pathway [7].

See also

References

  1. 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 Boven K-H, Fejtl M, Möller A, Nisch W, Stett A. On Micro-Electrode Array Revival. In: Baudry M, Taketani M, eds. Advances in Network Electrophysiology Using Multi-Electrode Arrays. New York: Springer Press; 2006: 24-37.
  2. Buitenweg JR, Rutten WL, and Marani E. 2003. Geometry-based finite element modeling of the electrical contact between a cultured neuron and a microelectrode. IEEE Trans Biomed Eng. 50: 501-509.
  3. 3.0 3.1 3.2 Potter SM. 2001. Distributed processing in cultured neuronal networks. Prog Brain Res 130: 49-62.
  4. 4.0 4.1 4.2 4.3 Pine J. A History of MEA Development. In: Baudry M, Taketani M, eds. Advances in Network Electrophysiology Using Multi-Electrode Arrays. New York: Springer Press; 2006:3-23.
  5. Buisson B, Heuschkel MO,Steidl EM, Wirth C. Development of 3-D Multi-Electrode Arrays for Use with Acute Tissue Slices. In: Baudry M, Taketani M, eds. Advances in Network Electrophysiology Using Multi-Electrode Arrays. New York: Springer Press; 2006:69-111.
  6. Thiebaud P, deRooij NF, Koudelka-Hep M, Stoppini L. 1997. Microelectrode arrays for electrophysiological monitoring of hippocampal organotypic slice cultures. IEEE Trans Biomed Eng. 44: 1159-63.
  7. 7.0 7.1 7.2 Cheung KC. 2007. Implantable microscale neural interfaces. Biomedical Microdevices 9: 923-38
  8. Thomas CA, Springer PA, Loeb GE, Berwald-Netter Y, Okun LM. 1972. A miniature microelectrode array to monitor the bioelectric activity of cultured cells. Exp Cell Res. 74: 61-66.
  9. Eytan D, Minerbi A, Ziv NE, Marom S. 2004. Dopamine-induced dispersion of correlations between action potentials in networks of cortical neurons. J Neurophysiol. 92(3): 1817-24.
  10. Segev R, Berry II, MJ. 2003. Recording from all of the ganglion cells in the retina. Soc Neurosci Abstr. 264: 11.
  11. Granados-Fuentes D, Saxena MT, Prolo LM, Aton SJ, Herzog ED. 2004. Olfactory bulb neurons express functional, entrainable circadian rhythms. Eur J Neurosci. 19(4): 898-906.
  12. Angelides KJ, Elmer LW, Loftus D, Elson E. 1988. Distribution and lateral mobility of voltage-dependent sodium channels in neurons. J Cell Biol. 106: 1911-25.
  13. Whitson J, Kubota D, Shimono K, Jia Y, Taketani M. Multi-Electrode Arrays: Enhancing Traditional Methods and Enabling Network Physiology. In: Baudry M, Taketani M, eds. Advances in Network Electrophysiology Using Multi-Electrode Arrays. New York: Spring Press; 2006: 38-68
  14. Biran R, Martin DC, Tresco PA. 2005. Neuronal cell loss accompanies the brain tissue response to chronically implanted silicon microelectrode arrays. Experimental Neurology 195: 115-26
  15. McConnell GC, Rees HD, Levey AI, Gross RG, Bellamkonda RV. 2008. Chronic electrodes induce a local, neurodegenerative state: Implications for chronic recording reliability. Society for Neuroscience, Washington, D.C
  16. He W, McConnell GC, Bellamkonda RV. 2006. Nanoscale laminin coating modulates cortical scarring response around implanted silicon microelectrode arrays. Journal of Neural Engineering 3: 316-26
  17. Gopal KV, Gross GW. Emerging Histotypic Properties of Cultured Neuronal Networks. In: Baudry M, Taketani M, eds. Advances in Network Electrophysiology Using Multi-Electrode Arrays. New York: Springer Press; 2006:193-214.
  18. Xia Y and Gross GW. 2003. Histotypic electrophysiological responses of cultured neuronal networks to ethanol. Alcohol 30: 167-74.
  19. DeMarse TB, Wagenaar DA, Blau AW, Potter SM. 2001. The Neurally Controlled Animat: Biological Brains Acting with Simulated Bodies. Autonomous Robots 11: 305-10.
  20. Potter, SM, Madhavan, R and DeMarse, TB. 2003. Long-term bidirectional neuron interfaces for robotic control, and in vitro learning studies. Proc. 25th IEEE EMBS Annual Meeting.
  21. Marks P. 2008. Rise of the rat-brained robots. New Scientist 2669.
  22. Colgin, L.L., Kramar, E.A., Gall, C.M., and Lynch, G. (2003). Septal modulation of excitatory transmission in hippocampus. J Neurophysiol. 90: 2358-2366.
  23. Breit S, Schulz JB, Benabid AL. 2004. Deep Brain Stimulation. Cell Tissue Research 318: 275-288.
  24. Warwick, K, Gasson, M, Hutt, B, Goodhew, I, Kyberd, P, Andrews, B, Teddy, P and Shad, A:“The Application of Implant Technology for Cybernetic Systems”, Archives of Neurology, 60(10), pp1369-1373, 2003
  25. Schwartz AB. 2004. Cortical Neural Prosthetics. Annual Review of Neuroscience 27: 487-507.


Template:WH Template:WS