Patch clamp

Last updated
A bacterial spheroplast patched with a glass pipette Patchclamp Spheroplast1.jpg
A bacterial spheroplast patched with a glass pipette
A patch clamp recording of current reveals transitions between two conductance states of a single ion channel: closed (at top) and open (at bottom). Single channel.png
A patch clamp recording of current reveals transitions between two conductance states of a single ion channel: closed (at top) and open (at bottom).

The patch clamp technique is a laboratory technique in electrophysiology used to study ionic currents in individual isolated living cells, tissue sections, or patches of cell membrane. The technique is especially useful in the study of excitable cells such as neurons, cardiomyocytes, muscle fibers, and pancreatic beta cells, and can also be applied to the study of bacterial ion channels in specially prepared giant spheroplasts.

Contents

Patch clamping can be performed using the voltage clamp technique. In this case, the voltage across the cell membrane is controlled by the experimenter and the resulting currents are recorded. Alternatively, the current clamp technique can be used. In this case, the current passing across the membrane is controlled by the experimenter and the resulting changes in voltage are recorded, generally in the form of action potentials.

Erwin Neher and Bert Sakmann developed the patch clamp in the late 1970s and early 1980s. This discovery made it possible to record the currents of single ion channel molecules for the first time, which improved understanding of the involvement of channels in fundamental cell processes such as action potentials and nerve activity. Neher and Sakmann received the Nobel Prize in Physiology or Medicine in 1991 for this work. [1]

Basic technique

Set-up

Classical patch clamp setup, with microscope, antivibration table, and micromanipulators Patch Clamp Rig classic.jpg
Classical patch clamp setup, with microscope, antivibration table, and micromanipulators

During a patch clamp recording, a hollow glass tube known as a micropipette or patch pipette filled with an electrolyte solution and a recording electrode connected to an amplifier is brought into contact with the membrane of an isolated cell. Another electrode is placed in a bath surrounding the cell or tissue as a reference ground electrode. An electrical circuit can be formed between the recording and reference electrode with the cell of interest in between.

Schematic depiction of a pipette puller device used to prepare micropipettes for patch clamp and other recordings Pipette Puller-en.svg
Schematic depiction of a pipette puller device used to prepare micropipettes for patch clamp and other recordings
Circuit formed during whole-cell or perforated patch clamp Patch clamp schematic 1.png
Circuit formed during whole-cell or perforated patch clamp

The solution filling the patch pipette might match the ionic composition of the bath solution, as in the case of cell-attached recording, or match the cytoplasm, for whole-cell recording. The solution in the bath solution may match the physiological extracellular solution, the cytoplasm, or be entirely non-physiological, depending on the experiment to be performed. The researcher can also change the content of the bath solution (or less commonly the pipette solution) by adding ions or drugs to study the ion channels under different conditions.

Depending on what the researcher is trying to measure, the diameter of the pipette tip used may vary, but it is usually in the micrometer range. [2] This small size is used to enclose a cell membrane surface area or "patch" that often contains just one or a few ion channel molecules. [3] This type of electrode is distinct from the "sharp microelectrode" used to puncture cells in traditional intracellular recordings, in that it is sealed onto the surface of the cell membrane, rather than inserted through it.

Typical equipment used during classical patch clamp recording Patch clamp schematic 2.png
Typical equipment used during classical patch clamp recording

In some experiments, the micropipette tip is heated in a microforge to produce a smooth surface that assists in forming a high resistance seal with the cell membrane. To obtain this high resistance seal, the micropipette is pressed against a cell membrane and suction is applied. A portion of the cell membrane is suctioned into the pipette, creating an omega-shaped area of membrane which, if formed properly, creates a resistance in the 10–100 gigaohms range, called a "gigaohm seal" or "gigaseal". [3] The high resistance of this seal makes it possible to isolate electronically the currents measured across the membrane patch with little competing noise, as well as providing some mechanical stability to the recording. [4]

Recording

Patch clamp of a nerve cell within a slice of brain tissue. The pipette in the photograph has been marked with a slight blue color. WholeCellPatchClamp-03.jpg
Patch clamp of a nerve cell within a slice of brain tissue. The pipette in the photograph has been marked with a slight blue color.

Many patch clamp amplifiers do not use true voltage clamp circuitry, but instead are differential amplifiers that use the bath electrode to set the zero current (ground) level. This allows a researcher to keep the voltage constant while observing changes in current. To make these recordings, the patch pipette is compared to the ground electrode. Current is then injected into the system to maintain a constant, set voltage. The current that is needed to clamp the voltage is opposite in sign and equal in magnitude to the current through the membrane. [3]

Alternatively, the cell can be current clamped in whole-cell mode, keeping current constant while observing changes in membrane voltage. [5]


Tissue sectioning

Accurate tissue sectioning with compresstome vibratome or microtomes is essential, in addition to patch clamp methods. By supplying thin, uniform tissue slices, these devices provide optimal electrode implantation. To prepare tissues for patch clamp studies in a way that ensures accurate and dependable recordings, researchers can select between using vibratomes for softer tissues and microtomes for tougher structures. [6] Leica Biosystems, Precisionary Instruments, Carl Zeiss AG are the notable producer of these devices.

Variations

Diagram showing variations of the patch clamp technique Patchmodes.svg
Diagram showing variations of the patch clamp technique

Several variations of the basic technique can be applied, depending on what the researcher wants to study. The inside-out and outside-out techniques are called "excised patch" techniques, because the patch is excised (removed) from the main body of the cell. Cell-attached and both excised patch techniques are used to study the behavior of individual ion channels in the section of membrane attached to the electrode.

Whole-cell patch and perforated patch allow the researcher to study the electrical behavior of the entire cell, instead of single channel currents. The whole-cell patch, which enables low-resistance electrical access to the inside of a cell, has now largely replaced high-resistance microelectrode recording techniques to record currents across the entire cell membrane.

Cell-attached patch

Cell-attached patch configuration Cell-Attached Patch Clamp.png
Cell-attached patch configuration

For this method, the pipette is sealed onto the cell membrane to obtain a gigaseal (a seal with electrical resistance on the order of a gigaohm), while ensuring that the cell membrane remains intact. This allows the recording of currents through single, or a few, ion channels contained in the patch of membrane captured by the pipette. By only attaching to the exterior of the cell membrane, there is very little disturbance of the cell structure. [3] Also, by not disrupting the interior of the cell, any intracellular mechanisms normally influencing the channel will still be able to function as they would physiologically. [7] Using this method it is also relatively easy to obtain the right configuration, and once obtained it is fairly stable. [8]

For ligand-gated ion channels or channels that are modulated by metabotropic receptors, the neurotransmitter or drug being studied is usually included in the pipette solution, where it can interact with what used to be the external surface of the membrane. The resulting channel activity can be attributed to the drug being used, although it is usually not possible to then change the drug concentration inside the pipette. The technique is thus limited to one point in a dose response curve per patch. Therefore, the dose response is accomplished using several cells and patches. However, voltage-gated ion channels can be clamped successively at different membrane potentials in a single patch. This results in channel activation as a function of voltage, and a complete I-V (current-voltage) curve can be established in only one patch. Another potential drawback of this technique is that, just as the intracellular pathways of the cell are not disturbed, they cannot be directly modified either. [8]

Inside-out patch

Inside-out patch configuration Inside-Out Patch Clamp.png
Inside-out patch configuration

In the inside-out method, a patch of the membrane is attached to the patch pipette, detached from the rest of the cell, and the cytosolic surface of the membrane is exposed to the external media, or bath. [9] One advantage of this method is that the experimenter has access to the intracellular surface of the membrane via the bath and can change the chemical composition of what the inside surface of the membrane is exposed to. This is useful when an experimenter wishes to manipulate the environment at the intracellular surface of single ion channels. For example, channels that are activated by intracellular ligands can then be studied through a range of ligand concentrations.

To achieve the inside-out configuration, the pipette is attached to the cell membrane as in the cell-attached mode, forming a gigaseal, and is then retracted to break off a patch of membrane from the rest of the cell. Pulling off a membrane patch often results initially in the formation of a vesicle of membrane in the pipette tip, because the ends of the patch membrane fuse together quickly after excision. The outer face of the vesicle must then be broken open to enter into inside-out mode; this may be done by briefly taking the membrane through the bath solution/air interface, by exposure to a low Ca2+ solution, or by momentarily making contact with a droplet of paraffin or a piece of cured silicone polymer. [10]

Whole-cell recording or whole-cell patch

Whole-cell patch configuration Whole Cell Patch Clamp Steps.png
Whole-cell patch configuration

Whole-cell recordings involve recording currents through multiple channels simultaneously, over a large region of the cell membrane. The electrode is left in place on the cell, as in cell-attached recordings, but more suction is applied to rupture the membrane patch, thus providing access from the interior of the pipette to the intracellular space of the cell. This provides a means to administer and study how treatments (e.g. drugs) can affect cells in real time. [11] Once the pipette is attached to the cell membrane, there are two methods of breaking the patch. The first is by applying more suction. The amount and duration of this suction depends on the type of cell and size of the pipette. The other method requires a large current pulse to be sent through the pipette. How much current is applied and the duration of the pulse also depend on the type of cell. [8] For some types of cells, it is convenient to apply both methods simultaneously to break the patch.

The advantage of whole-cell patch clamp recording over sharp electrode technique recording is that the larger opening at the tip of the patch clamp electrode provides lower resistance and thus better electrical access to the inside of the cell. [12] [11] A disadvantage of this technique is that because the volume of the electrode is larger than the volume of the cell, the soluble contents of the cell's interior will slowly be replaced by the contents of the electrode. This is referred to as the electrode "dialyzing" the cell's contents. [8] After a while, any properties of the cell that depend on soluble intracellular contents will be altered. The pipette solution used usually approximates the high-potassium environment of the interior of the cell to minimize any changes this may cause. There is often a period at the beginning of a whole-cell recording when one can take measurements before the cell has been dialyzed. [8]

Outside-out patch

Outside-out patch formation technique. In order: top-left, top-right, bottom-left, bottom-right Outside-Out Patch Clamp.png
Outside-out patch formation technique. In order: top-left, top-right, bottom-left, bottom-right

The name "outside-out" emphasizes both this technique's complementar­ity to the inside-out technique, and the fact that it places the external rather than intracellular surface of the cell membrane on the outside of the patch of membrane, in relation to the patch electrode. [7]

The formation of an outside-out patch begins with a whole-cell recording configuration. After the whole-cell configuration is formed, the electrode is slowly withdrawn from the cell, allowing a bulb of membrane to bleb out from the cell. When the electrode is pulled far enough away, this bleb will detach from the cell and reform as a convex membrane on the end of the electrode (like a ball open at the electrode tip), with the original outside of the membrane facing outward from the electrode. [7] As the image at the right shows, this means that the fluid inside the pipette will be simulating the intracellular fluid, while a researcher is free to move the pipette and the bleb with its channels to another bath of solution. While multiple channels can exist in a bleb of membrane, single channel recordings are also possible in this conformation if the bleb of detached membrane is small and only contains one channel. [13]

Outside-out patching gives the experimenter the opportunity to examine the properties of an ion channel when it is isolated from the cell and exposed successively to different solutions on the extracellular surface of the membrane. The experimenter can perfuse the same patch with a variety of solutions in a relatively short amount of time, and if the channel is activated by a neurotransmitter or drug from the extracellular face, a dose-response curve can then be obtained. [14] This ability to measure current through exactly the same piece of membrane in different solutions is the distinct advantage of the outside-out patch relative to the cell-attached method. On the other hand, it is more difficult to accomplish. The longer formation process involves more steps that could fail and results in a lower frequency of usable patches.

Perforated patch

Perforated patch technique Perforated Patch Clamp.png
Perforated patch technique

This variation of the patch clamp method is very similar to the whole-cell configuration. The main difference lies in the fact that when the experimenter forms the gigaohm seal, suction is not used to rupture the patch membrane. Instead, the electrode solution contains small amounts of an antifungal or antibiotic agent, such as amphothericin-B, nystatin, or gramicidin, which diffuses into the membrane patch and forms small pores in the membrane, providing electrical access to the cell interior. [15] When comparing the whole-cell and perforated patch methods, one can think of the whole-cell patch as an open door, in which there is complete exchange between molecules in the pipette solution and the cytoplasm. The perforated patch can be likened to a screen door that only allows the exchange of certain molecules from the pipette solution to the cytoplasm of the cell.

Advantages of the perforated patch method, relative to whole-cell recordings, include the properties of the antibiotic pores, that allow equilibration only of small monovalent ions between the patch pipette and the cytosol, but not of larger molecules that cannot permeate through the pores. This property maintains endogenous levels of divalent ions such as Ca2+ and signaling molecules such as cAMP. Consequently, one can have recordings of the entire cell, as in whole-cell patch clamping, while retaining most intracellular signaling mechanisms, as in cell-attached recordings. As a result, there is reduced current rundown, and stable perforated patch recordings can last longer than one hour. [15] Disadvantages include a higher access resistance, relative to whole-cell, due to the partial membrane occupying the tip of the electrode. This may decrease current resolution and increase recording noise. It can also take a significant amount of time for the antibiotic to perforate the membrane (about 15 minutes for amphothericin-B, and even longer for gramicidin and nystatin). The membrane under the electrode tip is weakened by the perforations formed by the antibiotic and can rupture. If the patch ruptures, the recording is then in whole-cell mode, with antibiotic contaminating the inside of the cell. [15]

Loose patch

Loose patch clamp technique Loose Patch Clamp.png
Loose patch clamp technique

A loose patch clamp is different from the other techniques discussed here in that it employs a loose seal (low electrical resistance) rather than the tight gigaseal used in the conventional technique. This technique was used as early as the year 1961, as described in a paper by Strickholm on the impedance of a muscle cell's surface, [16] but received little attention until being brought up again and given a name by Almers, Stanfield, and Stühmer in 1982, [17] after patch clamp had been established as a major tool of electrophysiology.

To achieve a loose patch clamp on a cell membrane, the pipette is moved slowly towards the cell, until the electrical resistance of the contact between the cell and the pipette increases to a few times greater resistance than that of the electrode alone. The closer the pipette gets to the membrane, the greater the resistance of the pipette tip becomes, but if too close a seal is formed, and it could become difficult to remove the pipette without damaging the cell. For the loose patch technique, the pipette does not get close enough to the membrane to form a gigaseal or a permanent connection, nor to pierce the cell membrane. [18] The cell membrane stays intact, and the lack of a tight seal creates a small gap through which ions can pass outside the cell without entering the pipette.

A significant advantage of the loose seal is that the pipette that is used can be repeatedly removed from the membrane after recording, and the membrane will remain intact. This allows repeated measurements in a variety of locations on the same cell without destroying the integrity of the membrane. This flexibility has been especially useful to researchers for studying muscle cells as they contract under real physiological conditions, obtaining recordings quickly, and doing so without resorting to drastic measures to stop the muscle fibers from contracting. [17] A major disadvantage is that the resistance between the pipette and the membrane is greatly reduced, allowing current to leak through the seal, and significantly reducing the resolution of small currents. This leakage can be partially corrected for, however, which offers the opportunity to compare and contrast recordings made from different areas on the cell of interest. Given this, it has been estimated that the loose patch technique can resolve currents smaller than 1 mA/cm2. [18]

Patch-Seq

A combination of cellular imaging, RNA sequencing and patch clamp this method is used to fully characterize neurons across multiple modalities. [19] As neural tissues are one of the most transcriptomically diverse populations of cells, classifying neurons into cell types in order to understand the circuits they form is a major challenge for neuroscientists. Combining classical classification methods with single cell RNA-sequencing post-hoc has proved to be difficult and slow. By combining multiple data modalities such as electrophysiology, sequencing and microscopy, Patch-seq allows for neurons to be characterized in multiple ways simultaneously. It currently suffers from low throughput relative to other sequencing methods mainly due to the manual labor involved in achieving a successful patch-clamp recording on a neuron. Investigations are currently underway to automate patch-clamp technology which will improve the throughput of patch-seq as well. [20]

Automatic patch clamping

Automated patch clamp systems have been developed in order to collect large amounts of data inexpensively in a shorter period of time. Such systems typically include a single-use microfluidic device, either an injection molded or a polydimethylsiloxane (PDMS) cast chip, to capture a cell or cells, and an integrated electrode.

In one form of such an automated system, a pressure differential is used to force the cells being studied to be drawn towards the pipette opening until they form a gigaseal. Then, by briefly exposing the pipette tip to the atmosphere, the portion of the membrane protruding from the pipette bursts, and the membrane is now in the inside-out conformation, at the tip of the pipette. In a completely automated system, the pipette and the membrane patch can then be rapidly moved through a series of different test solutions, allowing different test compounds to be applied to the intracellular side of the membrane during recording. [20]

See also

Related Research Articles

<span class="mw-page-title-main">Ion channel</span> Pore-forming membrane protein

Ion channels are pore-forming membrane proteins that allow ions to pass through the channel pore. Their functions include establishing a resting membrane potential, shaping action potentials and other electrical signals by gating the flow of ions across the cell membrane, controlling the flow of ions across secretory and epithelial cells, and regulating cell volume. Ion channels are present in the membranes of all cells. Ion channels are one of the two classes of ionophoric proteins, the other being ion transporters.

<span class="mw-page-title-main">Electrophysiology</span> Study of the electrical properties of biological cells and tissues.

Electrophysiology is the branch of physiology that studies the electrical properties of biological cells and tissues. It involves measurements of voltage changes or electric current or manipulations on a wide variety of scales from single ion channel proteins to whole organs like the heart. In neuroscience, it includes measurements of the electrical activity of neurons, and, in particular, action potential activity. Recordings of large-scale electric signals from the nervous system, such as electroencephalography, may also be referred to as electrophysiological recordings. They are useful for electrodiagnosis and monitoring.

<span class="mw-page-title-main">Action potential</span> Neuron communication by electric impulses

An action potential occurs when the membrane potential of a specific cell rapidly rises and falls. This depolarization then causes adjacent locations to similarly depolarize. Action potentials occur in several types of animal cells, called excitable cells, which include neurons, muscle cells, and in some plant cells. Certain endocrine cells such as pancreatic beta cells, and certain cells of the anterior pituitary gland are also excitable cells.

<span class="mw-page-title-main">Excitatory postsynaptic potential</span> Process causing temporary increase in postsynaptic potential

In neuroscience, an excitatory postsynaptic potential (EPSP) is a postsynaptic potential that makes the postsynaptic neuron more likely to fire an action potential. This temporary depolarization of postsynaptic membrane potential, caused by the flow of positively charged ions into the postsynaptic cell, is a result of opening ligand-gated ion channels. These are the opposite of inhibitory postsynaptic potentials (IPSPs), which usually result from the flow of negative ions into the cell or positive ions out of the cell. EPSPs can also result from a decrease in outgoing positive charges, while IPSPs are sometimes caused by an increase in positive charge outflow. The flow of ions that causes an EPSP is an excitatory postsynaptic current (EPSC).

Hyperpolarization is a change in a cell's membrane potential that makes it more negative. It is the opposite of a depolarization. It inhibits action potentials by increasing the stimulus required to move the membrane potential to the action potential threshold.

<span class="mw-page-title-main">Membrane potential</span> Type of physical quantity

Membrane potential is the difference in electric potential between the interior and the exterior of a biological cell. That is, there is a difference in the energy required for electric charges to move from the internal to exterior cellular environments and vice versa, as long as there is no acquisition of kinetic energy or the production of radiation. The concentration gradients of the charges directly determine this energy requirement. For the exterior of the cell, typical values of membrane potential, normally given in units of milli volts and denoted as mV, range from –80 mV to –40 mV.

<span class="mw-page-title-main">Threshold potential</span> Critical potential value

In electrophysiology, the threshold potential is the critical level to which a membrane potential must be depolarized to initiate an action potential. In neuroscience, threshold potentials are necessary to regulate and propagate signaling in both the central nervous system (CNS) and the peripheral nervous system (PNS).

<span class="mw-page-title-main">Voltage clamp</span>

The voltage clamp is an experimental method used by electrophysiologists to measure the ion currents through the membranes of excitable cells, such as neurons, while holding the membrane voltage at a set level. A basic voltage clamp will iteratively measure the membrane potential, and then change the membrane potential (voltage) to a desired value by adding the necessary current. This "clamps" the cell membrane at a desired constant voltage, allowing the voltage clamp to record what currents are delivered. Because the currents applied to the cell must be equal to the current going across the cell membrane at the set voltage, the recorded currents indicate how the cell reacts to changes in membrane potential. Cell membranes of excitable cells contain many different kinds of ion channels, some of which are voltage-gated. The voltage clamp allows the membrane voltage to be manipulated independently of the ionic currents, allowing the current–voltage relationships of membrane channels to be studied.

<span class="mw-page-title-main">Bert Sakmann</span> German Nobel laureate

Bert Sakmann is a German cell physiologist. He shared the Nobel Prize in Physiology or Medicine with Erwin Neher in 1991 for their work on "the function of single ion channels in cells," and the invention of the patch clamp. Bert Sakmann was Professor at Heidelberg University and is an Emeritus Scientific Member of the Max Planck Institute for Medical Research in Heidelberg, Germany. Since 2008 he leads an emeritus research group at the Max Planck Institute of Neurobiology.

<span class="mw-page-title-main">Erwin Neher</span> German biophysicist and Nobel laureate

Erwin Neher is a German biophysicist, specializing in the field of cell physiology. For significant contribution in the field, in 1991 he was awarded, along with Bert Sakmann, the Nobel Prize in Physiology or Medicine for "their discoveries concerning the function of single ion channels in cells".

The sucrose gap technique is used to create a conduction block in nerve or muscle fibers. A high concentration of sucrose is applied to the extracellular space, which prevents the correct opening and closing of sodium and potassium channels, increasing resistance between two groups of cells. It was originally developed by Robert Stämpfli for recording action potentials in nerve fibers, and is particularly useful for measuring irreversible or highly variable pharmacological modifications of channel properties since untreated regions of membrane can be pulled into the node between the sucrose regions.

In neuroscience, single-unit recordings provide a method of measuring the electro-physiological responses of a single neuron using a microelectrode system. When a neuron generates an action potential, the signal propagates down the neuron as a current which flows in and out of the cell through excitable membrane regions in the soma and axon. A microelectrode is inserted into the brain, where it can record the rate of change in voltage with respect to time. These microelectrodes must be fine-tipped, impedance matching; they are primarily glass micro-pipettes, metal microelectrodes made of platinum, tungsten, iridium or even iridium oxide. Microelectrodes can be carefully placed close to the cell membrane, allowing the ability to record extracellularly.

<span class="mw-page-title-main">Cable theory</span> Mathematical model of a dendrite

In neuroscience, classical cable theory uses mathematical models to calculate the electric current along passive neurites, particularly the dendrites that receive synaptic inputs at different sites and times. Estimates are made by modeling dendrites and axons as cylinders composed of segments with capacitances and resistances combined in parallel. The capacitance of a neuronal fiber comes about because electrostatic forces are acting through the very thin lipid bilayer. The resistance in series along the fiber is due to the axoplasm's significant resistance to movement of electric charge.

<span class="mw-page-title-main">Scanning ion-conductance microscopy</span> Scanning probe microscopy technique that uses an electrode as the probe tip

Scanning ion-conductance microscopy (SICM) is a scanning probe microscopy technique that uses an electrode as the probe tip. SICM allows for the determination of the surface topography of micrometer and even nanometer-range structures in aqueous media conducting electrolytes. The samples can be hard or soft, are generally non-conducting, and the non-destructive nature of the measurement allows for the observation of living tissues and cells, and biological samples in general.

<span class="mw-page-title-main">Dendritic spike</span> Action potential generated in the dendrite of a neuron

In neurophysiology, a dendritic spike refers to an action potential generated in the dendrite of a neuron. Dendrites are branched extensions of a neuron. They receive electrical signals emitted from projecting neurons and transfer these signals to the cell body, or soma. Dendritic signaling has traditionally been viewed as a passive mode of electrical signaling. Unlike its axon counterpart which can generate signals through action potentials, dendrites were believed to only have the ability to propagate electrical signals by physical means: changes in conductance, length, cross sectional area, etc. However, the existence of dendritic spikes was proposed and demonstrated by W. Alden Spencer, Eric Kandel, Rodolfo Llinás and coworkers in the 1960s and a large body of evidence now makes it clear that dendrites are active neuronal structures. Dendrites contain voltage-gated ion channels giving them the ability to generate action potentials. Dendritic spikes have been recorded in numerous types of neurons in the brain and are thought to have great implications in neuronal communication, memory, and learning. They are one of the major factors in long-term potentiation.

<i>Pflügers Archiv: European Journal of Physiology</i> Academic journal

Pflügers Archiv: European Journal of Physiology is a peer-reviewed scientific journal in the field of physiology. A continuation of a journal founded in 1868 by the German physiologist, Eduard Friedrich Wilhelm Pflüger, Pflügers Archiv is the oldest physiological journal. Pflügers Archiv is currently published by Springer, with 11 issues per year.

<span class="mw-page-title-main">Slice preparation</span> Laboratory method involving brain slices

The slice preparation or brain slice is a laboratory technique in electrophysiology that allows the study of neurons from various brain regions in isolation from the rest of the brain, in an ex-vivo condition. Brain tissue is initially sliced via a tissue slicer then immersed in artificial cerebrospinal fluid (aCSF) for stimulation and/or recording. The technique allows for greater experimental control, through elimination of the effects of the rest of the brain on the circuit of interest, careful control of the physiological conditions through perfusion of substrates through the incubation fluid, to precise manipulation of neurotransmitter activity through perfusion of agonists and antagonists. However, the increase in control comes with a decrease in the ease with which the results can be applied to the whole neural system.

Automated patch clamping is beginning to replace manual patch clamping as a method to measure the electrical activity of individual cells. Different techniques are used to automate patch clamp recordings from cells in cell culture and in vivo. This work has been ongoing since the late 1990s by research labs and companies trying to reduce its complexity and cost of patch clamping manually. Patch clamping for a long time was considered an art form and is still very time consuming and tedious, especially in vivo. The automation techniques try to reduce user error and variability in obtaining quality electrophysiology recordings from single cells.

Single-Entity Electrochemistry (SEE) refers to the electroanalysis of an individual unit of interest. A unique feature of SEE is that it unifies multiple different branches of electrochemistry. Single-Entity Electrochemistry pushes the bounds of the field as it can measure entities on a scale of 100 microns to angstroms. Single-Entity Electrochemistry is important because it gives the ability to view how a single molecule, or cell, or "thing" affects the bulk response, and thus the chemistry that might have gone unknown otherwise. The ability to monitor the movement of one electron or ion from one unit to another is valuable, as many vital reactions and mechanisms undergo this process. Electrochemistry is well suited for this measurement due to its incredible sensitivity. Single-Entity Electrochemistry can be used to investigate nanoparticles, wires, vesicles, nanobubbles, nanotubes, cells, and viruses, and other small molecules and ions. Single-entity electrochemistry has been successfully used to determine the size distribution of particles as well as the number of particles present inside a vesicle or other similar structures

Patch-sequencing (patch-seq) is a modification of patch-clamp technique that combines electrophysiological, transcriptomic and morphological characterization of individual neurons. In this approach, the neuron's cytoplasm is collected and processed for RNAseq after electrophysiological recordings are performed on it. The cell is simultaneously filled with a dye that allows for subsequent morphological reconstruction.

References

  1. "The Nobel Prize in Physiology or Medicine 1991". nobelprize.org. Nobel Media AB. Retrieved November 8, 2014.
  2. Bannister, Niel (November 1, 2012). Langton, Phil (ed.). Essential Guide to Reading Biomedical Papers: Recognizing and Interpreting Best Practice. Wiley-Blackwell. doi:10.1002/9781118402184. ISBN   9781118402184.
  3. 1 2 3 4 Sakmann, B.; Neher, E. (1984). "Patch clamp techniques for studying ionic channels in excitable membranes". Annual Review of Physiology. 46: 455–472. doi:10.1146/annurev.ph.46.030184.002323. hdl: 21.11116/0000-0000-D552-3 . PMID   6143532.
  4. Sigworth, Fredrick J.; Neher, E. (October 2, 1980). "Single Na+ channel currents observed in cultured rat muscle cells". Nature. 287 (5781): 447–449. Bibcode:1980Natur.287..447S. doi:10.1038/287447a0. PMID   6253802. S2CID   4238010.
  5. Ellen Covey; Matt Carter (2015). Basic Electrophysiological Methods. Oxford University Press. pp. 22–. ISBN   978-0-19-993980-0.
  6. Sakmann, B.; Edwards, F.; Konnerth, A.; Takahashi, T. (1989). "Patch clamp techniques used for studying synaptic transmission in slices of mammalian brain". Quarterly Journal of Experimental Physiology (Cambridge, England). 74 (7): 1107–1118. doi: 10.1113/expphysiol.1989.sp003336 . hdl: 11858/00-001M-0000-002C-270A-9 . PMID   2560557.
  7. 1 2 3 Hamill OP, Marty A, Neher E, Sakmann B, Sigworth FJ.; Marty; Neher; Sakmann; Sigworth (1981). "Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches". Pflügers Archiv: European Journal of Physiology. 391 (2): 85–100. CiteSeerX   10.1.1.456.107 . doi:10.1007/BF00656997. PMID   6270629. S2CID   12014433.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  8. 1 2 3 4 5 Molleman, Areles (March 6, 2003). Patch Clamping: An Introductory Guide To Patch Clamp Electrophysiology. Wiley. doi:10.1002/0470856521. ISBN   9780470856529.
  9. Veitinger, Sophie (2011-11-09). "The Patch-Clamp Technique". Science Lab. Leica Microsystems. Retrieved November 10, 2014.
  10. Ogden, David; Stanfield, Peter. "Patch Clamp Techniques" (PDF). utdallas.edu. pp. 53–78. Retrieved November 11, 2014.
  11. 1 2 Segev, Amir; Garcia-Oscos, Francisco; Kourrich, Saïd (2016-06-15). "Whole-cell Patch-clamp Recordings in Brain Slices". Journal of Visualized Experiments (112): e54024. doi:10.3791/54024. ISSN   1940-087X. PMC   4927800 . PMID   27341060.
  12. Staley, K.J.; Otis, T. S.; Mody, I (May 1, 1992). "Membrane properties of dentate gyrus granule cells: comparison of sharp microelectrode and whole-cell recordings". Journal of Neurophysiology. 67 (5): 1346–1358. doi:10.1152/jn.1992.67.5.1346. PMID   1597717.
  13. Howe, JR; Cull-Candy, SG; Colquhoun, D (Jan 1991). "Currents through single glutamate receptor channels in outside-out patches from rat cerebellar granule cells". Journal of Physiology. 432 (1): 143–202. doi:10.1113/jphysiol.1991.sp018381. PMC   1181322 . PMID   1715916.
  14. von Beckerath, N; Adelsberger, H; Parzefall, F; Franke, C; Dudel, J (Apr 1995). "GABAergic inhibition of crayfish deep extensor abdominal muscle exhibits a steep dose-response relationship and a high degree of cooperativity". European Journal of Physiology. 429 (6): 781–788. doi:10.1007/bf00374801. PMID   7541524. S2CID   7824699.
  15. 1 2 3 Linley, John (2013). "Perforated Whole-Cell Patch-Clamp Recording". In Gamper, Nikita (ed.). Ion Channels. Methods in Molecular Biology. Vol. 998 (Second ed.). Humana Press. pp. 149–157. doi:10.1007/978-1-62703-351-0_11. ISBN   978-1-62703-351-0. PMID   23529427.
  16. Strickholm, A (1 Jul 1961). "Impedance of a Small Electrically Isolated Area of the Muscle Cell Surface". Journal of General Physiology. 44 (6): 1073–88. doi:10.1085/jgp.44.6.1073. PMC   2195146 . PMID   19873540.
  17. 1 2 Almers W, Stanfield PR, Stühmer W (1983). "Lateral distribution of sodium and potassium channels in frog skeletal muscle: measurements with a patch clamp method". Journal of Physiology. 336 (10): 261–284. doi:10.1113/jphysiol.1983.sp014580. PMC   1198969 . PMID   6308223.
  18. 1 2 Lupa, MT; Caldwell, JH (Nov 1991). "Effect of Agrin on the Distribution of Acetylcholine Receptors and Sodium Channels on Adult Skeletal Muscle Fibers in Culture". Journal of Cell Biology. 115 (3): 765–778. doi:10.1083/jcb.115.3.765. PMC   2289169 . PMID   1655812.
  19. Tripathy, Shreejoy J.; Toker, Lilah; Bomkamp, Claire; Mancarci, B. Ogan; Belmadani, Manuel; Pavlidis, Paul (2018-10-08). "Assessing Transcriptome Quality in Patch-Seq Datasets". Frontiers in Molecular Neuroscience. 11: 363. doi: 10.3389/fnmol.2018.00363 . ISSN   1662-5099. PMC   6187980 . PMID   30349457.
  20. 1 2 Bowlby, Mark; Merrill, Thomas; Vasilyev, Dmitry (2005). "Development of a Novel Automated Ion Channel Recording Method Using Inside-Out Whole-Cell Membranes". Journal of Biomolecular Screening. 10 (8): 806–813. doi: 10.1177/1087057105279481 . PMID   16234349.