Physical Address
304 North Cardinal St.
Dorchester Center, MA 02124
Because virtually every biological process is sensitive to changes in pH, acid–base homeostasis is of critical importance to cells and organisms, and has attracted considerable attention. Until relatively recently, acid–base homeostasis, for both clinicians and basic scientists, has been synonymous with pH regulation in the two most easily accessed compartments, blood and cerebrospinal fluid (CSF). The pH in these extracellular compartments (pH o ) is certainly important for organisms. For example, alterations in pH o may affect various extracellular biochemical reactions (e.g., hemostasis, complement fixation) and influence binding of various substances (e.g., hormones, metals, therapeutic agents) to plasma proteins or cell surface receptors. Moreover, certain ion channels as well as transporters that move solutes across cell membranes are sensitive to pH o changes. Nevertheless, the number of processes sensitive to changes of extracellular pH pales in comparison with the myriad processes sensitive to alterations in intracellular pH (pH i ). Thus, pH i homeostasis should be a matter of central importance not only for individual cells, but also for the organism composed of these cells.
Although cellular metabolism can modulate pH i , the regulation of pH i is the province of membrane proteins that transfer acid–base equivalents across the plasma membrane. In addition, transporters that carry acid–base equivalents across organellar membranes can transiently modify pH i or can participate in the buffering of cytoplasmic H + .
Since the last edition of this book, the field of pH i regulation has continued to advance in the following broad areas: (1) the multiplicity of transport pathways responsible for pH i regulation, and (2) molecular mechanisms by which these pathways operate.
Because virtually every biological process is sensitive to changes in pH (for reviews, see refs. ), acid–base homeostasis is of critical importance to cells and organisms, and has attracted considerable attention. Until relatively recently, acid–base homeostasis, for both clinicians and basic scientists, has been synonymous with pH regulation in the two most easily accessed compartments, blood and cerebrospinal fluid (CSF). The pH in these extracellular compartments (pH o ) is certainly important for organisms. For example, alterations in pH o may affect various extracellular biochemical reactions (e.g., hemostasis, complement fixation) and influence binding of various substances (e.g., hormones, metals, therapeutic agents) to plasma proteins or cell surface receptors. Moreover, certain ion channels as well as transporters that move solutes across cell membranes are sensitive to pH o changes. Nevertheless, the number of processes sensitive to changes of extracellular pH pales in comparison with the myriad processes sensitive to alterations in intracellular pH (pH i ). Thus, pH i homeostasis should be a matter of central importance not only for individual cells, but also for the organism composed of these cells.
Although cellular metabolism can modulate pH i , the regulation of pH i is the province of membrane proteins that transfer acid–base equivalents across the plasma membrane. In addition, transporters that carry acid–base equivalents across organellar membranes can transiently modify pH i or can participate in the buffering of cytoplasmic H + .
Since the last edition of this book, the field of pH i regulation has continued to advance in the following broad areas: (1) the multiplicity of transport pathways responsible for pH i regulation, and (2) molecular mechanisms by which these pathways operate.
Elucidating the multiplicity of transport pathways in pH i regulation: A common theme is that patterns of pH i regulation in particular cell types is characteristic of the cell and complex. As a result, without previous knowledge of a cell’s physiology, it may be impossible to predict its response to a particular maneuver. For example, switching the extracellular buffer from a non-CO 2 /
buffer to CO 2 /
usually causes an abrupt fall in pH i , due to influx of the highly permeant CO 2 . On the other hand, some cell membranes show no evidence whatsoever of being permeable to gases such as CO 2 or NH 3 . In fact, some of the CO 2 permeability of membranes may require “gas” channels. For cells with CO 2 -permeable membranes, the response to the initial CO 2 -induced acidification may—depending on the cell type and initial pH i —be a pH i recovery that is totally absent, partial, complete, or even excessive.
Underlying the diversity of pH i regulation is its complexity. Thus, a particular cell type may possess numerous plasma-membrane transporters that regulate pH i , each with its own unique properties. For example, five genes encode Na + -coupled
transporters, and these generally have multiple variants, each of which has a characteristic expression profile that depends on cell type and developmental stage, and susceptibility to regulatory mechanisms. Since the last edition of this book, investigators have described many new variants of transporters and have clarified the roles that they play in pH i regulation.
pH i regulation is also complex in that Na + –H + exchangers, Na + -coupled
transporters, and other acid–base transporters may be under the concerted control of humoral agents or other environmental influences, such as acidosis or hypertonicity (see ref. ). In the case of pH o changes, certain G-protein-coupled receptors can detect extracellular H + . In addition, charges in the basolateral concentration of [CO 2 ] and [
] per se—independent of basolateral pH—are powerful regulators of acid–base transport in the proximal tubule.
Given the complexity and diversity of patterns of pH i regulation, one must examine pH i physiology anew for each previously unexplored cell type.
Understanding the molecular mechanisms of pH i regulation : Since the last edition of this book, the field has seen further major advances in the molecular physiology of acid–base transporters. Many of these carriers are discussed in greater detail in Chapter 53 on Na + -coupled
transporters, Chapter 54 on anion exchangers, and refs. on Na + -H + exchangers. The reader may also consult a review on Na + -coupled
transporters.
Scope of this chapter : In this chapter, we shall first consider the methodologies available for making pH i measurements. We shall then examine the factors that contribute to changes in pH i : the thermodynamic forces acting on hydrogen ions (H + ) and other charged acids/bases; the permeation of the cell membrane by electrically neutral acids and bases; the buffering power of the intracellular fluid; and transporters that regulate pH i by moving H + , bicarbonate ions (
) and/or other weak acid/bases across the plasma membrane. Because space does not permit us to summarize the vast array of acid–base transporters, we will focus on transporters that play a major role in pH i regulation. Emerging from this discussion will be a model of pH i regulation. Finally, in light of this model, we will consider several factors that fundamentally alter pH i regulation: alterations in pH o ; temperature changes; metabolic inhibitors and hypoxia; cell shrinkage; hormones, growth factors, and oncogenes.
Of the techniques available for measuring pH i , we shall consider the four that are currently of greatest utility. These techniques differ from one another in terms of their theoretical foundations, the precise cellular properties that they measure, as well as in their accuracy and sensitivity.
In the mid 1990s, it was beginning to look like the use of pH microelectrodes might become a dying art, except for their restricted use for measuring pH i in large invertebrate cells. However, the emergence of the Xenopus oocyte expression system and the cloning of numerous acid–base transporters over the past couple of decades have breathed new life into an old technology.
When a pH electrode and an indifferent reference electrode are placed in a solution, the voltage difference between the electrodes (E X ) is linearly related to the solution pH (pH X ):
pH S is the pH of a standard solution, E S is the voltage difference in this standard, and F/(RT ln 10) is the theoretical slope (~58 mV per pH-unit change at 22°C) of the line relating pH to voltage. The actual slope is determined empirically.
Although the pH sensor can be any of several materials (e.g., platinum-hydrogen, antimony, tungsten, or a liquid membrane), the most reliable remains pH-sensitive glass. Glass has the advantages of long lifetime, long-term stability, as well as insensitivity to cations other than H + (at physiological pH), redox reactions, and various gases. Several styles of glass pH-sensitive microelectrodes are available, including Hinke's exposed-tip design, Roger Thomas' recessed-tip design, and Roger Thomas’ eccentric design. Glass electrodes are particularly well suited to measure pH i of relatively large cells (e.g., squid axons), or to measure pH o . However, because it is difficult to fabricate glass microelectrodes that are small enough to be used with small cells, glass electrodes generally have been abandoned in favor of electrodes with liquid-membrane sensors, which are easier to make, but have a much shorter lifetime, are less stable, and are sometimes sensitive to drugs and other parameters.
Ideally, the pH and reference microelectrodes must impale a single cell. An acceptable alternative may be to place them in two identical cells, or in cells that are electrically coupled. Another solution is to employ a double-barreled electrode. Unfortunately, for each barrel, there is a trade-off between tip size and electrode performance. The larger the two barrels of a double-barreled electrode, the better the electrode performance, but the greater the cell damage caused by impalement. The performance of liquid-membrane electrodes can be improved by using a concentric design in which a saline-filled pipette is threaded into the column of a liquid-membrane sensor, thereby reducing the overall longitudinal resistance between the sensor at the electrode’s tip and the electrical contact in the electrode’s barrel. Fedirko et al. have described a simplified approach for implementing this concentric design for H + - and Ca 2+ -selective microelectrodes, permitting rapid measurements of extracellular pH and Ca 2+ transients in rat hippocampal brain slices.
pH-sensitive electrodes, particularly those with glass sensors, remain the method of choice for monitoring pH i in relatively large cells. Not only are these electrodes highly sensitive (<0.01 pH units), they are probably more precise and accurate than dyes. Moreover, the microelectrodes report pH and cell voltage simultaneously, and in real time. The “pH i ” that they report is almost certainly that of the bulk cytoplasm (i.e., the fluid in direct contact with the plasma membrane), uncontaminated by the pH of organelles. Disadvantages of using microelectrodes for measuring pH i include the time and skill required for making and using them. In addition, one can only use conventional microelectrodes on cells large enough to sustain the impalement.
Although small mammalian cells generally do not easily tolerate impalement, pH-sensitive microelectrodes placed near the extracellular side of the plasma membrane of a mammalian cell can be used to measure H + fluxes due to acid–base transporter activity. Using an elegant “self-referencing” pH microelectrode technique with Chinese hamster ovary (CHO) fibroblasts, Fuster et al. measured the continual extracellular H + gradent when a cell-attached patch pipette was repetitively positioned close to and away from an extracellular pH electrode. H + fluxes due to altered Na + –H + exchanger activity were quantitated from changes in the extracellular H + gradient elicited by altering the cytoplasm via pipette perfusion. Electrodes with relatively large tips (15–20 μm) have been used to monitor surface pH transients in oocytes and glean important information about the permeability to CO 2 and NH 3 in the study of gas channels.
Cell membranes are generally far more permeable to neutral molecules than to charged ones of similar shape and size. Thus, if a cell is exposed to a monoprotic weak acid HA (
), the neutral molecule rapidly enters the cell ( Figure 52.1 ). Assuming for the moment that A − cannot penetrate the membrane, the entry of HA continues until HA is in equilibrium across the cell membrane, that is, when the concentration of HA inside the cell ([HA] i ) is the same as that outside ([HA] o ). Because entering HA dissociates to H + and A − , the equilibration of HA across the membrane is necessarily accompanied by a fall in pH i . This principle underlies the pH i changes caused by neutral weak acids and bases that we will discuss below in the Section on “Effects of Weak Acids and Bases on pH i ”. Provided that the dissociation constant (K=[H + ]×[A − ]/[HA]) is the same both inside and outside the cell, then, at equilibrium,
Because the transmembrane distribution ratios of A − and H + are inversely related, we can use Equation 52.2 to compute pH i .
The weak acid most commonly used to calculate pH i is 5,5-dimethyl-2,4-oxazolidinedione (DMO), employed as 14 C-DMO; benzoic acid has also been used. Regardless of the weak acid used, the amount of radioactivity in the intra- or extracellular fluid is proportional to the total concentration of the probe ([A − ]+[HA]). pH i is computed from the following equation, which must include the concentrations of the total probe because it is impossible to measure [A − ] directly:
One can use a similar approach to compute pH i from the distribution of a permeant weak base, such as methylamine. Reviews by Waddell and Butler and by Roos and Boron contain more detailed descriptions and discuss potential difficulties of the weak-acid/base method.
The major advantages of the weak-acid/base method include its technical simplicity and applicability to even very small cells. The parameter actually measured is not the pH of the cytoplasm, but rather a volume–weighted mean pH of all intracellular compartments in which the weak acid or base is distributed. The practical sensitivity of this approach is 0.03–0.05 pH units—considerably less than the microelectrode technique. The major disadvantage of the weak-acid/base approach is that continuous pH i measurements are not possible.
Molecules with an absorbance, fluorescence excitation, and/or fluorescence emission spectrum sensitive to pH may be convenient probes for measuring pH i . When exposed to light, pH-sensitive dye molecules may absorb some of the light as electrons make the transition to a higher-energy state. From the intensities of incident light (I o ) and the light transmitted through the solution (I), we can compute the absorbance (A):
According to the Beer-Lambert law, A at a particular wavelength (A λ ) is proportional to both the length of the light’s path through the solution ( l ) and the concentration of the dye (C):
where the proportionality constant ε λ,pH is the dye’s extinction coefficient at a particular wavelength and pH. The wavelength dependence of ε λ,pH defines the shape of the absorbance spectrum, and the pH dependence of ε λ,pH defines how this shape is affected by changes in pH. However, A λ depends not only upon pH, but also upon l and C. Although l and C are extremely difficult to ascertain in a live cell, we can obtain the absorbance data at two wavelengths (λ 1 and λ 2 ), and compute the absorbance ratio (A λ1 /A λ2 ):
Thus, because the l and C terms cancel out, the absorbance ratio in Equation 52.6 , depends only on pH. By choosing two wavelengths such that ε λ1,pH /ε λ2,pH varies considerably with pH, one can obtain sensitive pH measurements. Experimenters using fluorescein derivatives typically use the peak absorbance wavelength (~510 nm), and the isosbestic wavelength (~440 nm), where ε is insensitive to pH changes. Monitoring dye absorbance at the isosbestic wavelength is attractive because one can determine the extent of dye loss during the experiment, assuming l is constant.
An advantage of absorbance for measuring pH i is that it tends to be extremely stable and sensitive. On the other hand, because absorbance is proportional to l and C, a relatively high intracellular dye concentration is required, even for thick preparations (e.g., renal tubules).
After absorbing a photon, most molecules return to the ground state by gradually losing energy through a series of random collisions with other molecules (see ref. ). Some dyes, however, can lose a quantum of energy from an excited singlet state by emitting a photon (i.e., fluorescing). The intensity of emitted fluorescent light (I emit ) can be measured with a photomultiplier tube (e.g., see ref. ) or an intensified [CCD] television camera, which provides imaging data (e.g., see ref. ). At most wavelengths, I emit is sensitive to pH, but at all wavelengths, I emit is sensitive to dye concentration as well as other parameters (e.g., position of cell in the incident light beam). Therefore, one usually employs a ratio technique to generate a parameter more uniquely related to pH. With the fluorescence-excitation ratio approach, commonly used with fluorescein dyes such as BCECF, one alternately excites at two wavelengths while monitoring I emit at one wavelength. With the fluorescence-emission ratio approach, commonly used with some rhodamine dyes such as SNARF-1, one excites at one wavelength, while monitoring I emit simultaneously at two wavelengths. By analogy with the absorbance ratio approach, one chooses three wavelengths to optimize the pH sensitivity of the ratio. A strength of the fluorescence-emission ratio approach is that the simultaneous capture of I emit at two wavelengths improves temporal resolution (see below).
Fluorescence measurements offer the advantage of being extremely sensitive. Thus, it is possible to quantitate the fluorescence from small amounts of dye, even in a microdomain within a single cell. In addition, one can use fluorescence with two-dimensional imaging, confocal microscopy, and multi-photon microscopy. Because fluorescence is more sensitive than absorbance to the environment of the dye molecule, fluorescence measurements are in principle more prone to artifact.
Excessive illumination can photolyse (i.e., bleach) dye molecules, causing a progressive decrease in the concentration of native dye. Among the photolysis products may be free radicals that react with cellular components and injure the cell (“photodynamic damage”). Excessive illumination can also cause a time-dependent shift in the apparent intracellular calibration curve of the dye, presumably due to bleaching-induced generation of dye-related products with spectral characteristics slightly different than those of the parent compound.
Bleaching can be particularly problematic during experiments on single cells, where the number of dye molecules and thus the number of emitted photons is low. The problem is that the limiting signal-to-noise ratio (S/N) is proportional to the square root of the number of photons measured. One can increase the number of emitted photons by increasing the intensity of the excitation light source, but at the expense of exacerbating the bleaching and photodynamic damage. The damage can be minimized either by illuminating continuously with low-intensity light for longer, or by limiting the duty cycle of high-intensity exciting light. Thus, one must sometimes trade off pH i resolution (i.e., S/N) against time resolution. In imaging experiments, where a cell may be represented by thousands of picture elements (pixels), the photon emission rate per pixel is exceedingly low. Here, one must trade off pH i resolution, time resolution, and spatial resolution (i.e., the number of pixels that must be grouped to compute pH i values).
More recent advances in fluorescence microscopy have helped to minimize the predicament of good pH i vs. spatial/temporal resolution. For example, two-photon excitation laser scanning microscopy is a relatively new imaging technique that has allowed investigators to reduce photobleaching of ion-sensitive fluorescent dyes, even while acquiring high spatial and temporal resolution recordings. Multiphoton imaging technology is based on the general quantum principle that a molecule can exhibit fluorescence after absorbing two photons simultaneously when excited by high-intensity light of twice the wavelength necessary for single-photon absorption. Excitation at the higher wavelengths (and thus lower energies) reduces overall photobleaching of the dye. Two-photon microscopy has been particularly useful in measuring intracellular Ca 2+ transients with excellent spatial resolution in tissues and cellular microdomains (see ). The technique has also been used with pH-sensitive dyes. For example, intracellular pH has been measured with BCECF in microdomains of the epidermis, and with SNARF-4 in villous enterocytes in vivo .
Because dyes interact with cytoplasmic components, the spectroscopic properties of an intracellular dye differ, sometimes markedly, from those of the same dye examined in a cuvette. Therefore, intracellular dye calibration is essential. One calibration approach is to use a second, independent method for measuring pH i in either the same cell, or another cell under similar conditions. For example, simultaneous measurements with a pH-sensitive microelectrode have confirmed that the absorbance indicator dimethylcarboxyfluorescein in salamander proximal-tubule cells, and the fluorescence indicator BCECF in leech glial astrocytes yield reasonable values.
The most popular approach has been to monitor intracellular absorbance or fluorescence while using the high-[K + ] o /nigericin technique to clamp pH i to predetermined values. Nigericin is a carboxylic ionophore that exchanges K + (and to a lesser extent Na + ) for H + across cell membranes. If one is successful in choosing [K + ] o to match [K + ] i , then pH i should equal pH o . Thus, by altering pH o , one can measure the dye’s spectral properties over a range of pH i values. A detailed calibration spanning a wide pH i range can be obtained for each experiment. Alternately, one can perform the detailed calibration on one set of cells, and routinely perform only a single-point calibration. Potential problems with the high-[K + ] o /nigericin technique have been discussed in some detail.
In using nigericin-containing solutions for dye calibration, one must be careful to cleanse the perfusion system completely after each calibration procedure. Even trace amounts of nigericin can interfere with the assessment of pH i -regulating mechanisms, by mimicking a K + –H + exchanger and by increasing “background acid loading a
a We define “acid loading” as any process that causes pH i to fall. Examples include the uptake of H + , the loss of HCO 3 − , or the metabolic production of H + .
” through nigericin-mediated exchange of internal K + for external H + .
A novel calibration technique, originally proposed for microelectrodes by Szatkowski and Thomas, but
applied to dyes by Eisner et al., involves sequentially exposing the cell to a permeant weak acid and weak base. As discussed below, the size of a pH i change elicited by a weak acid/base depends on the initial pH i . It is thus possible to compute the initial pH i from the magnitude of the change of the pH i indicator. Eisner et al. have extended this approach by introducing an elegant null-point technique, in which one experimentally determines a combination of weak-acid and weak-base concentrations that produces no change in the measured fluorescence. The null-point approach can be particularly useful in assessing the validity of other dye-calibration procedures, particularly the high-[K + ] o /nigericin technique.
The most popular dye for fluorescence measurements of pH i is the fluorescein derivative BCECF, which has four negatively-charged carboxylate groups and a phenolic -OH moiety that is titrated by pH changes. The dye can be directly loaded into large cells (e.g., Xenopus oocytes) with an injection pipette, or into small mammalian cells by diffusion from a patch pipette during whole-cell recordings. However, BCECF is usually loaded into cells as an uncharged acetoxymethylester (AM) precursor that easily permeates most plasma membranes. Intracellular esterases hydrolyze BCECF-AM to yield four or five formaldehyde molecules for each charged BCECF molecule trapped inside the cell. The time required for dye loading can vary greatly among cell types, from 1 minute to tens of minutes. Less conventional methods for dye loading—including scrape loading, osmotic lysis, and electroporation—are examined in more detail in Giuliano and Taylor. BCPCF, with carboxypropyl groups, is a derivative of BCECF that can be used as a dual-emission pH indicator.
Both pyranine-based and rhodamine-based dyes have also been used for monitoring pH i . Rhodamine dyes such as the seminaphthorhodafluor (SNARF) dyes are excited and emit at longer wavelengths than the fluorescein derivatives. An advantage of SNARF is that its fluorescence-emission spectrum is pH sensitive, in addition to its absorbance and fluorescence-excitation spectra. Thus, the dye can be used in the dual-emission mode, and is therefore more useful than fluorescein derivatives for confocal microscopy, in which one typically excites at only one wavelength. Dual-emission dyes also enable the user to sample more frequently, and to avoid the delay in alternating between two excitation wavelengths. SNARF-1 can be used simultaneously with Fura-2 for monitoring both pH i and [Ca 2+ ] i . In general however, one should be cautious in using multiple ion-sensitive dyes simultaneously to avoid quenching artifacts.
pH-sensitive fluorescent dyes with lower pK values are available (e.g., from Life Technologies™/Molecular Probes®) for pH measurements in acidic organelles such as lysosomes. Some of these dyes include the weak-base LysoSensor probes, the aminorhodamine dye pHrodo, and the fluorinated fluorescein dye Oregon Green. These probes can be targeted to appropriate organelles by a cell’s own endocytic pathway.
Investigators have also developed ingenious methods for targeting pH-sensitive probes to organelles. As reviewed by Maxfield and Yamashiro, pH-sensitive indicators can be targeted to the endocytic pathway following pinocytosis or receptor-mediated endocytosis. Kim et al. have measured Golgi pH with either a rhodamine- or fluorescein-labeled β subunit of verotoxin, which accumulates in the Golgi complex after receptor-mediated endocytosis and retrograde transport. Grinstein’s group has also measured pH in the endoplasmic reticulum and the trans -Golgi network by creating chimeric proteins with organelle-specific retrieval signals, and subsequently tagging them with pH-sensitive fluorophores before internalization. A similar approach has been used to measure the pH of recycling endosomes. The Machen group has examined pH regulation in the secretory pathway by targeting biotin-labelled pH probes to organelles expressing avidin-chimera proteins. Seksek et al. have used a different technique to measure the trans -Golgi pH of fibroblasts. They injected the cells with 70-nm liposomes containing membrane-impermeable pH-sensitive fluorophores; the liposomes then fused with the trans -Golgi. As described in the next section, investigators have also measured the pH of organelles by fusing pH-sensitive green fluorescent protein (GFP) variants or associated biosensors to organellar localization signals.
Another fluorophore that has become useful in the pH field is GFP, which is a natural product of the jellyfish Aequorea victoria , and typically used to label proteins expressed in cells. Several researchers have engineered GFP mutants such as pHluorins that are sensitive to pH changes in the physiological range, and can be targeted to the cytosol or organelles. Wild-type or mutant fluorescent proteins such as GFP and YFP can also be fused together to create ratiometric pH biosensors, and subsequently target the biosensors to organelles such as the mitochondia. Tantama et al. have recently engineered—from the red fluorescent protein mKeima—a pH-sensitive, dual-excitation/ratiometric variant called pHRed. This variant would be particularly useful during simultaneous measurements requiring the use of other more blue-shifted fluorescent probes. Moreover, upon two-photon, single-wavelength excitation, pHRed exhibits modest pH-sensitive changes in fluorescence lifetime that could be used to obtain rough estimates of pH i .
Compared to more traditional pH indicators, pH-sensitive GFP mutants are advantageous in displaying a low rate of photobleaching while remaining trapped inside of cells. pHluorins linked to markers of synaptic vesicles (synaptopHluorins) undergo marked changes in fluorescence upon exocytosis at synaptic terminals and have therefore been used to characterize synaptic vesicle cycling associated with presynaptic activity. Such vesicular cycling has been examined in transgenic mice expressing synaptopHluorin. It is intriguing to speculate on the potential applicability of pH-sensitive GFPs to examine cellular pH physiology. Indeed, pHluorin constructs have been used to measure pH i of fungi, plants, and yeast, as well as transiently transfected HEK293 cells and astrocytes (Liu and Bevensee, unpublished). One could perform in vivo pH i measurements on targeted cells that are induced by a specific promoter to express a pH-sensitive GFP. Metzger et al. successfully measured pH i in cerebellar slices from a generated transgenic mouse expressing a pH- and chloride-sensitive yellow-green variant of GFP (EYFP) under the control of a neuronal potassium channel promoter.
One can exploit differential dye loading in preparations to target dyes to specific cells or locations. For example, intercalated cells in the cortical collecting tubule of the kidney incorporate BCECF-AM much more rapidly than the neighboring principal cells. In a somewhat different application, Chu et al. found that mouse colonic crypt cells exclude SNARF, which can then be used to measure extracellular pH changes in the intact epithelium. Also, a probenecid-sensitive organic-anion transporter can extrude BCECF from some cells including thyroid cells, and epithelial kidney and intestinal cells. In an elegant study, Harris et al. used BCECF to measure the pH of the lateral intracellular space (LIS) of MDCK cell monolayers by first loading the cells with BCECF, and subsequently allowing organic transporters to move the dye into the LIS.
BCECF generally stains the cytoplasm rather uniformly, although in some cells the nuclear region is more intense than peripheral areas. Working on Ehrlich ascites tumor cells, Thomas et al. found that 6-carboxyfluorescein is confined to the cytoplasm, but fluorescein is distributed in both the mitochondria and cytoplasm. Furthermore, Slayman et al. found that BCECF can accumulate in vacuoles of the fungus Neurospora when the cells are exposed to BCECF-AM.
One can evaluate a dye’s intracellular compartmentalization by monitoring the fluorescence loss elicited by selectively permeabilizing different compartments with detergents. For example, most of the BCECF loaded into rat hippocampal CA1 neurons appears to be in the cytoplasm because 0.01% saponin reduces the fluorescence signal by ~96%. Both absorbance- and fluorescence-derived estimates of pH i can be influenced by dye in compartments other than the cytoplasm, depending upon each compartment's volume, dye concentration and pK, and pH.
Certain atomic nuclei, among them 31 P and 19 F, possess a quantum mechanical property termed “spin,” and behave as tiny bar magnets with magnetic moments. When an atomic nucleus of this type is placed in an external magnetic field, the magnetic moment precesses with a characteristic frequency about the axis of the applied field. The nucleus can be excited to a high-energy state by irradiating it with an oscillating magnetic field of the same frequency (i.e., resonance frequency) as the precession frequency. The resonance frequency depends not only on the identity of the atomic nucleus (e.g., 31 P), but also on its chemical environment, which influences the strength of the magnetic field at the nucleus. Thus, the resonance frequencies for 31 P in
and
are slightly different because of the different chemical environments of the 31 P. Because the exchange rate of 31 P between individual
and
ions is very rapid, nuclear magnetic resonance (NMR) detects only a single inorganic phosphate peak, the location of which depends on [
]/[
]. Because the dependence of this ratio on pH is described by a modified pH-titration curve (
), the position of the inorganic phosphate peak is a good index of pH i .
A major advantage of 31 P-NMR is that, in addition to providing nearly continuous measurements of pH i , it can also be used to monitor levels of a variety of phosphorus-containing compounds, such as ATP. The pH i value derived from NMR measurements is predominantly the pH of the cytoplasm. However, the inorganic phosphate peak for the cytosol may overlap with the inorganic-phosphate peaks from other intracellular compartments, as well as the extracellular space. The mitochondrial inorganic phosphate peak can in some cases be resolved from that of the cytoplasmic peak. NMR measurements have been made on whole organs, whole small animals, and human limbs.
pH i can also be measured by NMR with 19 F-labeled probes having pK values in the physiological range. As discussed by Deutsch, 19 F-labeled probes are advantageous over 31 P in that background signals are low, and the fluorinated probes are highly visible and sensitive to the environment. Commonly used probes such as fluoroanilines, derivatives of fluoroisobutyric acid, and fluorinated pyridoxins are generally introduced into cells as methyl esters, similar to the approach used for pH-sensitive dyes. Aside from cost and the need for technical expertise, the major disadvantage of NMR is its relatively low sensitivity. Thus, a considerable mass of 31 P or 19 F is required for detection, precluding the use of the technique with single cells. For reviews, see refs .
As reviewed by Gallagher et al., related approaches for estimating pH i include a comparison of 13 C signals for CO 2 and
, the use of 89 Y-labeled compounds, and pH-sensitive exchange of 1 H 2 O between bulk water and water bound to a gadolinium complex.
Until the 1930s, it was generally assumed that hydrogen ions were in electrochemical equilibrium across the cell membrane, as defined by the Nernst equation:
Here, V m is the voltage difference across the plasma membrane (the unit is the volt), R is the universal gas content (8.31 joules • °K −1 • equivalent −1 ), T is absolute temperature, F is Faraday's constant (96,486 coulombs • equivalent −1 ), and the subscripts o and i refer to extracellular and intracellular, respectively. In terms of pH, the foregoing equation becomes
assuming a temperature of 22°C. Thus, if H + were in equilibrium, then pH i would be one unit lower than pH o for each −58.5 mV of membrane potential. However, such a situation would present severe problems for the cell, inasmuch as V m changes would shift the equilibrium distribution of H + , and thus alter pH i and pH i -sensitive processes. It was not until the mid-1930s that Fenn and colleagues, in experiments on frog skeletal muscle, demonstrated that pH i is higher than expected for H + to be in electrochemical equilibrium. The authors, in effect, estimated V m from the ratio [K + ] o /[K + ] i , and measured pH i using the weak-acid method.
Nowadays, both V m and pH i can be measured directly, and for nearly all cells studied, pH i is well above the equilibrium pH i . In vertebrate skeletal muscle, for example, V m is ~−90 mV and pH i is ~7.1 3 . Given a pH o of 7.4, Equation 52.8 predicts an equilibrium pH i of ~5.9, far lower than the actual pH i . Rather than calculating equilibrium pH, one could equally well compute equilibrium potential for H + , that is, the membrane voltage required for H + to be in equilibrium across the membrane (E H ). In skeletal-muscle example introduced above (pH i =7.1, pH o =7.4), the E H computed from Equation 52.7 or Equation 52.8 is ~−18 mV. Because the actual V m (−90 mV) is more negative than E H , H + is drawn into the relatively negative cell, driven by an electrochemical gradient of 90–18 or 72 mV. Thus, there is a substantial electrochemical gradient (1.2 pH units or 72 mV) favoring the passive influx of H + . If the cell membrane were permeable to H + , then the resultant H + influx would represent a chronic intracellular acid load (i.e., an acid load imposed on the cell for an indefinite period) that would tend to lower pH i .
In the above discussion, we made no assumptions about the mechanism of the hypothetical influx of H + . It had generally been thought that H + flux across the plasma membrane occurs via nonspecific pathways. However, as extensively reviewed by DeCoursey, voltage-activated, Zn 2+ -inhibited H + currents were first described in snail neurons, and have subsequently been characterized in numerous other cell types, including epithelial cells (e.g., kidney cells), connective tissue, skeletal muscle, lymphocytes, macrophages, granulocytes, and microglia. Ramsey et al. and Sasaki et al. independently identified the first cDNA encoding a Zn 2+ -sensitive, voltage-gated proton channel, termed H v 1 (also known as VSOP). As first shown by Meech and Thomas, the channel is closed at normal V m levels, when the H + electrochemical gradient favors the passive influx of H + . However, strong depolarization not only reverses the H + gradient—now favoring the passive efflux of H + —but also opens the channels. Thus, as reviewed by Capasso et al., as long as the cell is highly depolarized, this channel functions as an H + -efflux pathway that can contribute to acid extrusion, defined below in our discussion of pH i regulation. Murphy et al., used the
-prepulse technique to acid load rat alveolar epithelial cells, and computed the acid-extrusion rate from the subsequent pH i recovery. They observed similar Zn 2+ -sensitive acid extrusion, regardless of whether the cells were depolarized with high-K + solutions, and concluded that H v 1 channels must have been contributing to the pH i recovery. However, it is not clear whether—under conditions of ‘normal’ membrane voltage—the H + electrochemical gradient would be outward and the H v 1 channels would be open. In the absence of data showing that the Zn 2+ effects are reduce/eliminated by the knockdown/knockout of H v 1, one must entertain the hypothesis that Zn 2+ had nonspecific effects and that H v 1 did not contribute to pH i regulation.
The depolarization threshold for channel activation is lowered by increasing the pH o -to-pH i gradient. Although the single-channel conductance is low for H v 1, high protein expression in cells account for macroscopic H + currents that can exceed K + currents.
It has become clear that H v 1 is functionally coupled to NADPH oxidases (NOX), and plays a critical role in removing the protons formed as byproducts during the generation of
; for reviews, see refs . This H v 1 activity is important during neutrophil phagocytosis, signaling in immune cells, as well as in spermatozoa activation (see ref. ).
In general, the passive fluxes of ionic weak acids (e.g.,
) and bases (e.g.,
) also impose a chronic intracellular acid load. As noted earlier (see Equation 52.2 ), whenever a neutral weak acid (HA) is equilibrated across the cell membrane ([HA] i =[HA] o ), the transmembrane distribution ratio for H + is the reciprocal of the distribution ratio for the anionic conjugate weak base (A − ). Consider, for example, the CO 2 /
buffer system. If (1) [CO 2 ] i equals [CO 2 ] o , (2) CO 2 is in equilibrium with H + and
both inside and outside the cell, and (3) the equilibrium constant is the same inside and outside the cell, then:
That is, the electrochemical gradient for
(and that for any other monovalent anion that can be described by a similar equation) is equal to but opposite to the gradient for H + . Inasmuch as H + normally leaks into cells,
and other anionic weak bases tend to leak out, also decreasing pH i . In crayfish muscle, the transmitter GABA opens Cl − channels that are also permeable to
. The resultant
efflux can reduce pH i by as much as 0.4. A similar GABA A -activated
conductance is present in cells from turtle cerebellum and rat hippocampus. This GABA A -activated
conductance can be inhibited by antagonists of GABA A receptors such as picrotoxin. In a similar fashion,
movement through glycine-activated Cl − channels alters pH i . Other anionic weak bases that seem to penetrate at least some cell membranes include the DMO anion, formate, propionate, and salicylate.
We can make a similar analysis for cationic weak acids (
). If the electrically neutral conjugate weak base (B) is equilibrated across the cell membrane (i.e., [B] i =[B] o ), if the equilibrium
holds on both sides of the membrane, and if the equilibrium constant is the same on both sides of the membrane, then:
That is, the electrochemical gradient for HB + is in the same direction as that for H + . Thus, similar to H + , cationic weak acids such as
tend to enter the cell and produce a chronic intracellular acid load.
The preceding analysis shows that, under the conditions that normally prevail in most cells (e.g., pH i ≅7.1, V m ≅−60), the electrochemical gradients affecting H + and charged monovalent weak acids/bases (provided the neutral species equilibrates across the membrane) generally favor fluxes that would lower pH i . Note that the above thermodynamic analysis addresses only the net direction of these passive fluxes, and does not address the rate of intracellular acid loading. These passive fluxes are depicted for a model cell in Figure 52.2 . Additional acid-loading mechanisms include carrier-mediated transport of H + into or
out of cells, as well as metabolism that generates acidic byproducts. Acid-loading mechanisms such as passive fluxes, carrier-mediated transport, and metabolism are “chronic” because they act continuously to lower pH i . Maintaining a normal pH i requires that acid loading be matched by a comparable—and continuous—extrusion of acid. By definition, this acid extrusion b
b We define “acid extrusion” as any process that causes pH i to rise. Examples include the efflux of H + and the uptake of HCO 3 − .
must be an active (i.e., energy-requiring) process. Acid extrusion can be accomplished either by the active uptake of alkali (e.g., OH − or
) and/or the active removal of acid (e.g., H + ).
In the section above on “Distribution of Weak Acids and Bases” for measuring pH i , we discussed the flux of neutral weak acids and bases across the cell membrane. In general, solutions containing the weak acid HA always contain the conjugate weak base A − . However, because membranes are usually far more permeable to HA than to A − , HA fluxes generally have a greater effect on pH i .
When a cell is exposed to a neutral weak acid, HA enters and dissociates into A − and H + , thereby causing pH i to fall:
This movement of the neutral weak acid (or a neutral weak base such as NH 3 , discussed below) has been termed non-ionic diffusion. This process continues until [HA] i equals [HA] o . Subsequently, there should be no further change in pH i , provided there is no flux of A − and no change in the rates of acid–base transporter activity. Although CO 2 is often regarded as a weak acid, it is not an acid at all. Only after reacting with H 2 O does CO 2 yield the true weak acid, H 2 CO 3 , which then dissociates (as described above for the idealized weak acid HA):
These two reactions can be combined into a single one with an overall apparent equilibrium constant (
[H + ]/[CO 2 ]). The enzyme carbonic anhydrase (CA) greatly accelerates the formation of
from CO 2 by catalyzing the reaction:
which generates H + by virtue of consuming OH − . The mechanism by which CA catalyzes this reaction is examined in more detail by Liljas et al. In the absence of CA, the reaction can also occur, particularly under alkaline conditions (i.e., pH >8).
In the example of Figure 52.3 , a cell is successively exposed to solutions equilibrated with 1, 2 and 5% CO 2 (constant [
] o =10 mM). Each time, CO 2 produces a rapid and sustained fall of pH i . The period during the CO 2 exposure in which pH i is relatively stable is termed the plateau phase . The magnitude of the CO 2 -induced acidification is inversely related to the intracellular buffering power (β; see below). The magnitude of the pH i decrease also increases with [CO 2 ] o . The degree of dissociation is governed by the relationship between pH i and
, which, in logarithmic form, is the familiar Henderson-Hasselbalch equation:
where
is ~6.1 at 37°C. Therefore, if pH i is 7.1, 10 molecules of incoming CO 2 dissociate into H + plus
for each molecule of CO 2 that remains CO 2 . If the initial pH i is only 6.1, then this ratio falls to 1:1, and fewer incoming CO 2 molecules dissociate. There are two practical consequences of this rule. First, the lower the initial pH i , the smaller the magnitude of acidification elicited by the subsequent exposure to the CO 2 . Second, as one raises [CO 2 ] o by fixed increments, the magnitude of the acidification is not proportional to the successive [CO 2 ] o increment. Thus, in Figure 52.3 , the ΔpH i produced by 2% CO 2 is less than twice as large as that produced by 1% CO 2 , and that produced by 5% CO 2 is substantially less than five times that produced by 1% CO 2 .
Although investigators have long believed that all gases penetrate all membranes simply by dissolving in the membrane lipid, Waisbren et al. demonstrated the first membrane with negligible permeability to a dissolved gas (i.e., CO 2 and NH 3 ). Moreover, Nakhoul et al. and Cooper and Boron identified the first gas channel—AQP1, which is permeable to CO 2 . Later work showed that AQP1 is also permeable to NH 3 and that rhesus (Rh) proteins can be permeable to both CO 2 and NH 3 . In human erythrocytes, AQP1 and the Rh complex, together, are responsible for ~90% of the CO 2 permeability. Thus, at least two families of proteins can function as gas channels. Furthermore, just as ion channels display ion selectivity, the AQPs and Rh proteins display gas selectivity. Based on CO 2 - or NH 3 -mediated changes in the surface pH of oocytes expressing AQP or Rh-family members, the sequence of CO 2 /NH 3 selectivity is AQP4≅AQP5>AQP1>AmtB>RhAG.
Weak acids such as acetic acid, lactic acid, and DMO can also elicit intracellular acidifications, which have [HA] o and pH i dependencies similar to those described for CO 2 . Aside from the dissolved gases discussed in the previous paragraph, many non-volatile neutral weak acids and bases may move through plasma membranes—at least in part—via protein pathways (e.g., channels, transporters) in addition to any traffic through the lipid phase of the membrane. In a general sense, all such fluxes can be thought of as non-ionic diffusion, regardless of the mechanism. CO 2 is unusual, however, in that [CO 2 ] o is easily controlled, independent of [
] o , either by the experimenter (who can equilibrate solutions with a known CO 2 mixture) or by the intact organism (which can alter its external respiration). Of course, the price one pays for varying [CO 2 ] at constant [
] is that—at equilibrium—pH must vary as well, as predicted by Equation 52.14 . In the laboratory, it is possible to overcome this limitation by using out-of-equilibrium CO 2 /
solutions (see below) to change only one of the three parameters (pH, CO 2 , or
) at a time. For nonvolatile weak acids, the experimenter can only manipulate the total concentration of extracellular buffer ([TA] o =[HA] o +[A − ] o ) and pH. When [TA] o is fixed, changes in pH o produce reciprocal alterations in [HA] o and [A − ] o , so that the magnitude of the HA-induced pH i decrease is very pH o sensitive.
In the preceding discussion, we assumed that the movement of the uncharged weak acid HA was the only factor affecting pH i . In other words, we excluded the possibilities that: (1) the charged species A − can traverse the cell membrane and significantly affect pH i , and (2) acid–base transporters as well as metabolic processes that generate acid–base equivalents can alter the overall balance between acid extrusion/loading. In the absence of these complicating effects, applying and withdrawing HA/A − would produce the pH i changes illustrated in Figure 52.4b .
What would happen if A − , as well as HA, were able to cross the membrane passively? At the instant a cell is exposed to a solution containing HA and A − , the electrochemical gradient for A − would be inward (assuming that the initial [A − ] i were sufficiently low). Initially then, both HA and A − would enter, the HA tending to lower pH i , and the A − tending to raise pH i . A significant permeability of the membrane to A − would reduce the initial rate of the HA-induced acidification. Eventually, however, the generation of A − from incoming HA, as well as the influx of A − , would raise [A − ] i to such a level that A − would tend to exit the cell passively, as described earlier. In addition, a transporter might move A − out of the cell; for example, a Cl −
exchanger normally moves
out of the cell. Subsequently, the efflux of A − would lower pH i , both during the initial HA-induced acidification and during the later plateau phase, as illustrated in Figure 52.4a . This A − efflux represents what might be termed a “semi-chronic” acid load, inasmuch as the pH i decrease would continue until [A − ] i was so low that A − was equilibrated across the membrane. Removing external HA/A − would elicit a rapid rise in pH i due to the efflux of HA from the cell. Note, however, that pH i would rise to a value less than the initial pH i . The magnitude of the shortfall would be directly related to the net acid loading due to A − efflux. The pH i changes illustrated in Figure 52.4a , presumably due to permeability of the membrane to the A − form of a weak acid, have been observed in snail neurons exposed to salicylic acid.
Acid-extrusion mechanisms can also influence the pH i changes of a cell exposed to HA/A − . In practice, the effects described below are due to one or more acid-extruders that outstrip acid loaders. These transporters are described in more detail later, as are their effects on the dynamics of pH i regulation. We introduce these transporters here only to complete our consideration of pH i transients elicited by HA/A − . Consider a cell that is permeable to HA, but not to A − , and that has an acid-extrusion mechanism that is unaffected by A − (e.g., Na + H + exchange). Applying HA/A − would acutely acid load the cell, due to the influx of HA, and subsequent formation of A − and H + . However, an acid-extruder stimulated by the pH i decrease would remove this acute acid load and return pH i to exactly its starting value ( Figure 52.4C ). Note that removing HA/A − causes pH i to overshoot its initial value by a substantial amount. The magnitude of the overshoot is directly related to the net amount of acid extruded during the HA/A − exposure. This pattern of pH i changes, including a complete recovery of pH i during the plateau phase, is observed in snail neurons acid loaded by exposure to CO 2 /
. Other examples have been reported in which the plateau-phase pH i recovery is incomplete.
As illustrated in Figure 52.4d , exposing a cell to HA/A − may cause the pH i during the plateau phase to rise to a value even greater than the initial pH i . For example, if the cell had an acid extruder that is stimulated either by HA or A − , then the influx of HA would initially acidify the cell. However, the stimulated acid extruder would not merely return pH i to its initial value, as in Figure 52.4c , but increase pH i beyond the initial value, as in Figure 52.4d . We can provide two such examples in which A − stimulates acid extrusion and thereby elicits pH i changes similar to those in Figure 52.4d : (1) exposing renal mesangial cells to CO 2 /
stimulates a Na + -driven
uptake mechanism, and (2) adding acetic acid/acetate to the lumen of a rabbit S3 proximal tubule stimulates a Na + /acetate cotransporter. HA can also stimulate acid extrusion. In the rabbit proximal tubule, CO 2 per se appears to stimulate Na + H + exchange and H + pumping.
Solutions containing the neutral weak base B also contain its conjugate weak acid HB + . Because membranes are usually far more permeable to B than to HB + , B fluxes tend to have a greater effect on pH i .
When a cell is exposed to a neutral weak base, B enters and associates with H + to form HB + , causing pH i to rise:
This process continues until [B] i equals [B] o . Subsequently, there should be no further change in pH i , provided there is no flux of HB + and no change in the rates of acid–base transporter activity. Removing external B reverses the reaction as written in Equation 52.15 , and B passively diffuses out of the cell. pH i should then return to exactly its initial level. An example of the effect of the weak base NH 3 is shown in Figure 52.5 . Applying 20 mM total ammonium (i.e., [NH 3 ] o +[
] o ) causes pH i first to rise, and then stabilize (segment ab, pulse 1). Removing the NH 3 /
elicits a fall in pH i to a value somewhat lower than the initial one (compare points a and c). The reason for this small pH i undershoot will become clear below. As discussed above for HA-induced acidifications, the magnitude of the NH 3 -induced alkalinization depends upon intracellular buffering power, [NH 3 ] o , and the degree to which entering NH 3 is protonated. The last is governed by the relation:
where
(~9.2 at 22°C) is the acid dissociation constant. Thus, at the initial pH i of 7.3 (point a), approximately 99.4% of entering NH 3 is protonated to form
, whereas at a pH i of 7.8 (point b), only approximately 98% is protonated. The dependence of the protonation of NH 3 on the difference (
–pH i ) has two consequences, analogous to the two discussed earlier for CO 2 . First, the higher the initial pH i , the smaller the magnitude of the alkalinization elicited by the subsequent exposure to the
. Second, as one raises [NH 3 ] o by fixed increments, the magnitude of the alkalinization with each successive [NH 3 ] o increment does not increase in proportion. Thus, in Figure 52.5 , the alkalinization produced by 10 mM total
is less than twice as great as that produced by 5 mM total
.
In Figure 52.5 , the magnitude of the NH 4 + -induced alkalinization actually depends more on [NH 3 ] o than on [total
] o . Thus, the
-induced alkalinizations will be identical for solutions in which [total
] and pH o are varied reciprocally so as to keep [NH 3 ] o constant. Neutral weak bases such as lidocaine, procaine, and methylamine also elicit intracellular alkalinizations that have [B] o and pH i dependencies similar to those described for NH 3 .
As noted above, the bacterial protein AmtB, an Rh homolog, functions as a gas channel permeable to NH 3 . In addition, aquaporins 1, 4, 5, and AQP8 are permeable to NH 3 .
It is possible to target the weak base NH 3 and selectively raise the pH of a specific intracellular compartment. For example, Carraro-Lacroix et al. raised the pH of lysosomes by first endocytotically targeting urease to the lysosome compartment, and then exposing the cells to urea. The urease—now in the lysosomes—catalyzed the hydrolysis of the urea to produce one molecule of CO 2 but two of the weak base NH 3 . Both because of the 2:1 stoichiometry of generated NH 3 :CO 2 and the low initial lysosomal pH, the NH 3 dominates and the net effect is the NH 3 -driven consumption of H + to form
, thereby selectively alkalinizing the lysosome.
When one exposes a cell to a weak base, the dominant effect on pH i generally reflects the influx of the highly permeant neutral weak base (e.g., NH 3 ). However, the flux of the conjugate weak acid (e.g.,
) may produce pH i decreases that are substantial or even dominant. The effects on pH i of applying and withdrawing NH 3 /
can be separated into four steps. When a cell is exposed to NH 3 /
, the rapid influx of NH 3 leads to an initial rise in pH i ( Figure 52.6a ). However, this alkalinization is actually blunted by the simultaneous, although smaller, influx of
. The most general mechanism of
influx is the passive diffusion of
, although the carrier-mediated transport of
into cells can be substantial. For example,
can enter cells by substituting for K + on the Na + K + pump and/or the Na + /K + /Cl − cotransporter. Although most of the entering
remains
, a small fraction (governed by the differences between
and pH i ) dissociates to form NH 3 and H + . When external NH 3 /NH 4 + is removed, the NH 4 + that previously entered, but failed to dissociate, now dissociates into NH 3 (which rapidly leaves the cell) and H + (which is trapped within the cell). As a consequence, the final pH i (point c in Figure 52.5 ) is slightly lower than the initial value (point a).
The acidifying effect of
is much more evident when the NH 3 /
exposure lasts beyond the initial influx ( Figure 52.6a ) and equilibration of NH 3 ( Figure 52.6b ). During an extended plateau phase, the continuing net influx of NH 4 + produces a plateau-phase acidification, as illustrated in Figure 52.6c . In cells with a Cl −
exchanger (e.g., sheep cardiac Purkinje fibers), this plateau-phase acidification may be augmented by efflux of
in exchange for Cl − . When the external NH 3 /NH 4 + is eventually withdrawn, the pH i undershoot is greatly exaggerated ( Figure 52.6d ). The magnitude of the undershoot reflects the previous influx of
. The longer the exposure to NH 3 /
, the more extensive the plateau-phase acidification, and the larger the undershoot. This undershoot represents an acute intracellular acid load. Indeed, the
-prepulse technique is widely used for acid loading cells in studies of pH i regulation.
If the
electrochemical gradient is reversed during the plateau phase, in a cell for which the dominant mechanism of
entry is a passive flux, then pH i rises during the plateau phase and the pH i undershoot is converted to a shortfall.
In the broadest sense, a pH buffer is any system that moderates the effects of an acid load by reversibly consuming H + , or of an alkali load by reversibly releasing H + . As we shall see, buffering in the intracellular fluid (e.g., the cytosol) is considerably more complex than in the blood and extracellular fluids. One can determine the buffering power (β) of either the extra- or intracellular fluid by applying an acute acid or alkali load, and measuring the resultant change in pH. β is defined as the amount of strong base (or strong acid) required to raise (or lower) pH by a given amount. Strictly speaking, the definition is given in differential notation. However, if the amount of base added (ΔB, given in mM) and the resultant pH increase (ΔpH) are sufficiently small, then β is approximated by:
This definition was originally proposed in slightly different form by Koppel and Spiro in 1914 and later modified to its present state independently by Michaelis and Van Slyke. Note that β is always a positive number; for a strong acid, ΔB and ΔpH are both negative. β is usually reported in mM per pH unit.
Among the several methods available for applying acid and alkali loads to the cytoplasm, the most direct is by injection of acid or alkali into the cell with a micropipette. Such an acute acid load produces an abrupt fall in pH i (see Figure 52.7 , segment ab), the extent of which is determined by the amount of acid injected, the intracellular buffering power, and the ability of pH i -regulatory mechanisms to blunt the pH i decrease. As discussed in later sections, these pH i -regulatory mechanisms can actively extrude acid from the cell and thereby return pH i to its initial value (segment b–f). If we were to calculate β blindly, then we would see that the apparent β depends critically on the time of the ΔpH i measurement following the acid load. For example, the apparent β is relatively low at point b and rises to infinity at point f, where ΔpH i is zero.
An approach for avoiding the above ambiguity is to define intracellular buffering power (β i ) under conditions in which we completely block pH i regulation. In this case, pH i would follow the trajectory indicated by ag in Figure 52.7 , the observed ΔpH i at “infinite” time would be relatively large, and the computed β would be smaller than any of the possibilities described above. There are three justifications for excluding acid–base transport processes from the definition of β i : (a) As we have just seen, excluding cellular transport processes makes the calculated β i time independent (provided we allow enough time for pH i to fall). (b) As we shall see later, transport of acid and/or base can generally be distinguished from buffering mechanisms by applying transport inhibitors or performing ion substitutions in the extracellular fluid. (c) Generally, these transport processes are not fully reversible. The buffering mechanisms included in our definition of β i fall into three categories: (a) weak acids and bases (i.e., chemical buffers), (b) biochemical reactions consuming or releasing H + , and (c) transport of acids or bases across organellar membranes. These mechanisms, which will be discussed in more detail later, can be justifiably grouped because there is presently no practical basis for distinguishing among them when examining cytosolic pH regulation. Furthermore, they proceed very rapidly compared to the time currently required to measure pH i .
Buffering mechanisms abate pH i changes produced by acute acid and alkali loads that otherwise could damage the cell. However, these mechanisms cannot prevent a change in pH i , only reduce its magnitude. Furthermore, buffering mechanisms cannot restore pH i to its initial value following an acid or alkali load. Such recoveries are brought about by transport of acid and/or base across the cell membrane. As an illustration, consider the response of the cell to an acute intracellular acid load, as in Figure 52.7 . Cellular buffers respond very rapidly, typically consuming more than 99.99% of the applied acid and limiting the initial fall in pH i (ag). Active transport mechanisms respond more slowly, extruding acid from the cell and returning pH i toward normal (b–f). During the transport-mediated pH i recovery, H + previously taken up by the buffering mechanisms is now released and extruded from the cell. By the time pH i returns to its initial level (f), the entire acid load has been extruded from the cell, and cellular buffering mechanisms have been returned to their pre-acid-loading state (a). Buffering mechanisms play no role in the steady state; in this case, acid loading exactly balances acid extrusion, and the state of cellular buffers is unchanged.
The total intracellular buffering power (β T ) is the sum of chemical (i.e., weak-acid/base equilibria), biochemical, and organellar buffering powers.
A weak base (B) of valence n reacts with H + to produce its conjugate weak acid (HB) of valence n+1:
This equilibrium is described by the relation:
where
is the apparent acid dissociation constant. This relation can also be expressed in logarithmic form, whereupon it takes the form of Equation 52.16 . Starting with the definition of β ( Equation 52.17 ) and the statement of chemical equilibrium ( Equation 52.19 ), one can express β as a function of [H + ],
, and total buffer concentration.
where 2.3 approximates ln(10), and [TB] is the concentration of total buffer (i.e., [TB]=[B n ]+[HB n+1 ]). By obtaining the derivative of β with respect to [H + ] (i.e., d β/ d [H + ]) and setting this equal to zero, one can show that β is maximal when [H + ] equals
. This maximal buffering power (β max ) is simply:
Although the pH of maximal buffering power is in general unique to the buffer, all buffers have the same molar buffering power (β max /[TB]), and the same dependence of β on the difference pH–
. If more than one buffer is present, then β T is simply the sum of the individual buffering powers, each computed from Equation 52.20 .
The foregoing equation describing β was derived by taking the partial derivative of [B n ] with respect to pH at a fixed [TB]. That is, the analysis applies only to buffers in a closed system, in which [TB] is constant. Closed-system buffers include a solution of a nonvolatile buffer (e.g., inorganic phosphate) in a beaker, a solution of a volatile buffer (e.g., CO 2 /
) in a capped syringe, and an impermeant buffer (e.g., a protein) in the cytoplasm of a cell. In such a closed system, β for CO 2 /
would be rather low at normal pH i values, because pH i (~7.0) is far higher than the
of 6.1. However, cells are generally not a closed system with respect to CO 2 /
. Rather, CO 2 usually freely exchanges with the extracellular fluid (ECF) so that the total intracellular CO 2 (i.e., [CO 2 ] i +[
] i ) can change appreciably during intracellular acid–base disturbances. For example, when H + is added to cytoplasm containing CO 2 /
, H + combines with
according to the reactions:
The cytoplasmic CO 2 so formed is lost to the ECF. Thus, it is [CO 2 ] i that remains constant, while [
] i and [total CO 2 ] i both fall. Because there is no buildup of CO 2 within the cell, the extent of the reaction in Equation 52.22 is limited only by the availability of
. The amount of H + absorbed (i.e., buffered) by the CO 2 /
system exactly equals the amount of
that disappears. That is:
where β CO 2 is the CO 2 buffering power. When this equation, in differential form, is combined with the Henderson-Hasselbalch equation, it can be shown that:
when [CO 2 ] is held constant. At very high pH values, when the equilibrium
cannot be ignored, the term 4.6 [
] must be added to the right-hand side of Equation 24 . This equation describes the buffering power of CO 2 /
in an open system, such as a solution in a beaker equilibrated with gaseous CO 2 . Because most cell membranes are highly permeable to CO 2 , this gas rapidly equilibrates across membranes and stabilizes [CO 2 ] i , provided the ECF behaves as an infinite reservoir for CO 2 . Thus, an isolated cell in vitro behaves as an open system for CO 2 /
. Because the intact organism has mechanisms (e.g., alveolar ventilation in higher vertebrates) for stabilizing [CO 2 ] in the ECF, CO 2 /
behaves as an open-system buffer in vivo , both in the extra- and intracellular spaces.
In an open system, the CO 2 /
buffer pair generally makes a substantial contribution to β T . For example, at a pH i of 7.1, the buffering power of all non-CO 2 buffers in rat renal mesangial cells is only about 10 mM, whereas the computed β CO 2 is nearly 29 mM when the cell is equilibrated with 5% CO 2 . Thus, β CO 2 accounts for nearly 75% of the β T of 39 mM/pH unit.
It is essential to distinguish clearly between open- and closed-system buffers of a cell. The cell is a closed system for intracellular buffers that do not readily cross the cell membrane (e.g., inorganic phosphate and the imidazole groups of proteins). Thus, these buffers are influenced by the attributes of the cell (e.g., volume, temperature, and metabolic state), and not those of the extracellular fluid. These are termed intrinsic buffers . Biochemical and organellar buffering mechanisms ( vide infra ) are also intrinsic, and the total intrinsic buffering power (β I ) is the sum of biochemical, organellar, and intrinsic chemical buffering powers.
The cell is an open system when one member of the buffer pair readily crosses the cell membrane. The buffering power of such a buffer pair is very sensitive to extracellular conditions. As noted earlier, β CO 2 is proportional to [
] i , which in turn is completely determined by pH i and the extracellular P CO 2 (assuming that CO 2 equilibrates across the cell membrane). The cell also behaves as an open system for buffers other than CO 2 /
. These are generally conjugate pairs of which one member is a small, neutral molecule. Examples include NH 3 /
and acetic acid/acetate. Regardless of whether the charged species is a cation or an anion, the intracellular buffering power of an open-system buffer pair is always 2.3×[charged species] i , provided the neutral species is equilibrated across the cell membrane. Thus, the intracellular buffering power of such a buffer pair is sensitive not only to the total amount of the buffer in the ECF (i.e., [TB] o ), but also to pH o . As with CO 2 /
, these other open-system buffers can sometimes be the dominant component of β T . For example, β I in squid axons is ~9 mM. When these axons are exposed to a pH o −7.7 NH 3 /
solution containing only 10 mM total ammonium, [
] i rises to 10 mM. The computed β NH 3 is 23 mM, nearly 70% of β T . Because the buffering power of such open-system buffers is so sensitive to factors external to the cell (e.g., pH o ), these buffers are termed extrinsic . The total chemical buffering power of a cell is the sum of the buffering powers of the individual conjugate acid/base buffer pairs, be they intrinsic (closed system) or extrinsic (open system).
Because certain biochemical reactions consume or release H + and are pH i sensitive, they can act as H + buffers. Examples include hydrolysis of ATP (which releases H + ) and hydrolysis of phosphocreatine (which consumes H + ). During the Cori cycle, the liver converts lactic acid into glucose, whereas skeletal muscle breaks down the glucose to produce ATP and more lactic acid. Also, H + buffering can arise when reactions induce a change in the
of an ionizable group. A classic example is the buffering reactions of oxygenated versus deoxygenated hemoglobin.
A well studied example of biochemical buffering are reactions involving intermediary metabolites in rat brain. Cells that are acutely acid loaded with an increase of P CO 2 in the ECF respond by reducing intracellular levels of several acid metabolic intermediates (i.e., the acids of lactate, pyruvate, citrate, α-ketoglutarate, malate, glutamate, and aspartate), while raising those of glucose and glucose-6-phosphate. These observations are consistent with the hypothesis that reducing pH i inhibits a step in the glycolytic pathway, possibly the phosphofructokinase reaction. Indeed, mouse muscle phosphofructokinase is markedly sensitive to pH changes ; reducing in-vitro pH from 7.2 to 7.1 produces more than a 90% inhibition. Rat-brain cells respond in the opposite fashion to acute intracellular alkali loads. Increases in pH i induced by lowering the P CO 2 in the ECF lead to increased levels of lactate and pyruvate. Thus, the biochemical machinery of these cells seems to respond appropriately to pH i changes, consuming H + in response to intracellular acid loads and releasing H + in response to alkali loads. The extent to which such biochemical reactions contribute to overall buffering power can be computed from changes in steady-state metabolite levels, provided that such changes are reciprocally linked to the production or consumption of neutral and/or readily diffusible molecules. For example, the consumption of one citrate molecule removes three H + , whereas malate removes two, and pyruvate, one. From data on rat brain, the buffering power of biochemical reactions amounts to about half that provided by all non-CO 2 physicochemical buffers. Biochemical buffering power can be defined and quantitated in a manner analogous to chemical ( vide supra ) and organellar ( vide infra ) buffering powers [see Equation 52.17 ]. The biochemical buffering reaction can be written as:
where R is the reactant, P is the product, n is the valence, and q is the H + stoichiometry ( q >0). The biochemical buffering power is thus:
As noted above, biochemical and organellar ( vide infra ) buffering mechanisms, in addition to closed-system chemical buffers, comprise the intrinsic buffers.
H + -transport systems have been identified or implicated in many intracellular organelles, including mitochodria, lysosomes, sarcoplasmic reticulum, Golgi network, endosomes, chromaffin granules, and zymogen granules. In addition, three of the cloned Na-H exchangers (NHE6, 7, and 9) are found in organellar membranes where they may contribute to organellar pH regulation. For reviews of vesicular H + pumps, see refs. It would be reasonable to expect decreases in cytoplasmic pH to stimulate H + uptake and/or inhibit H + extrusion by at least some organelles, and increases in pH i to produce the opposite effects. Such buffering would lead to a net transfer of H + into these organelles following intracellular acid loads, and a net movement of H + into the cytoplasm following alkali loads. A transfer of acid or base across organellar membranes would constitute a de facto buffer mechanism for the bulk intracellular fluid. Although the extent to which such hypothetical organellar buffering mechanisms contribute to β T is not established, it can be inferred from published data that changes in extraorganellar pH can produce at least small changes in the intraorganellar pH of mitochondria, lysosomes, and sarcoplasmic reticulum vesicles. Organellar buffering power can be defined and quantitated in a manner analogous to that for physicochemical and biochemical buffering [see Equation 52.17 ]. The organellar buffering reaction is written:
where the parentheses refer to organellar contents. Organellar buffering power is thus:
Organellar buffers, along with the biochemical and closed-system chemical buffers, comprise β I .
The easiest method for estimating β i is to homogenize a tissue sample and titrate the homogenate. The slope of the pH titration curve is the buffering power of the homogenate. However, some disadvantages limit the accuracy of this technique. First, homogenization may disrupt cellular organelles and thus obscure the organellar buffering contribution. Second, homogenization-induced changes in metabolism must be prevented, for example by quick-freezing the sample before homogenization, and then performing the titration either at low temperature or in the presence of an inhibitor such as fluoride. Blocking metabolism will reduce biochemical buffering, and reducing temperature generally increases
values, thereby altering chemical buffering.
The most straightforward approach for estimating β i in an intact cell is to microinject a known amount of acid or base into a cell and monitor the resultant change in pH i . The microinjection can be achieved either by iontophoresis or by pressure injection. This technique enables one to compute the total β i (i.e., sum of physicochemical, biochemical, and organellar buffering) of the cell. For small neurons incubated in a CO 2 -free solution, the iontophoresis and pressure-injection methods give β i values of 10.8 mM/pH unit and 10.3 mM/pH unit, respectively. Disadvantages of this direct in-vitro titration approach include the necessity of using cells that (1) are large enough to withstand the microelectrode impalements, and (2) have a sufficiently compact geometry to permit rapid equilibration of the injected acid or base throughout the cell.
An approach that, at least in principle, is applicable to cells of all shapes and sizes, is measurement of the pH i change produced by exposing the cell to a neutral weak acid HA (
), or weak base B (
). As discussed earlier, exposing a cell to a neutral weak acid leads to a decrease in pH i as HA enters the cell and partially dissociates into H + and A − . The entry and dissociation halt once [HA] i =[HA] o . The extent of the dissociation depends on the weak acid’s
, as well as the initial pH i . Virtually all of the released H + is consumed by intracellular buffers ( Figure 52.8A and B ); the remaining fraction is responsible for the accompanying pH i decrease. Although HA/A − is a buffer, it does not participate in buffering the acid load imposed by the intracellular dissociation of HA. Similarly, a B/BH + buffer system can not buffer changes in pH i caused by the entry (or exit) of B. Because the dissociation of HA leads to the formation of one A − for each H + , the magnitude of the intracellular acid load is Δ[A − ] i . This amount of added acid is the additive inverse of the amount of base added (ΔB) in the preceding definition of β:
Note that β i is the non-HA/A − buffering power of the cell. It is the sum of chemical, biochemical, and organellar buffering. ΔpH i is directly measured, and Δ[A − ] i is calculated from final and initial values of pH i and [HA] o . In the simplest case, the initial [A − ] i is zero, and final [A − ] i is given by:
assuming [HA]=[HA] i =[HA] o . A commonly used weak acid for measuring β i is CO 2 , for which:
β i is the non-CO 2 /
or intrinsic buffering power,
. Once
is known, the total intracellular buffering power (
) can be obtained by computing the open-system CO 2 buffering power:
where
, the CO 2 buffering power, is given by 2.3×[
] i , as outlined above.
There are as many as three major drawbacks of using the weak-acid method to measure β. First, [HA] i may not exactly equal [HA] o . In rapidly metabolizing cells for instance, [CO 2 ] i may exceed [CO 2 ] o . Second, the pH i decrease elicited by the entry of HA may stimulate acid-extrusion mechanisms (i.e., pH i -regulating systems described below). Thus, the actual ΔpH i may be smaller than if all acid–base transport had been blocked ( Figure 52.7 ). By underestimating ΔpH i , one will overestimate β i . This error should in principle be eliminated by blocking the pH i -regulating systems. One can also minimize the error by extrapolating the pH i -versus-time curve back to a point at which acid extrusion is judged to have had no effect on pH i . For example, in the hypothetical experiment of Figure 52.7 , one could extrapolate the pH i recovery (i.e., bcdef) back to the time of the acid load. A third potential drawback of the weak-acid method is that the weak acid could indirectly alter pH i by modifying cellular metabolism. Thus, it is critical that the weak acid affect pH i only by entering and dissociating into H + and A − . Otherwise, any other effect on pH i will lead to an error in the calculation of β i .
The intrinsic intracellular buffering power can also be determined by exposing a cell to a neutral weak base, B ( Figure 52.8B ). As described in the preceding section, exposing a cell to B leads to a pH i increase as most of the entering B combines with H + to form HB + . The entry of B and the alkali loading of the cell continues until [B] i =[B] o . Almost all of the entering H + that is converted to HB + is derived from cellular buffers. The minute amount that comes from the pool of free H + is responsible for the rise in pH i . Thus, the change in [HB + ] i is equivalent to the amount of strong base added to the intracellular fluid:
In the simplest case, the initial [HB + ] i is zero, and the final [HB + ] i is calculated from the statement of chemical equilibrium:
assuming that [B]=[B] i =[B] o . The most commonly used weak base is NH 3 :
[
] i is determined in a manner analogous to that outlined earlier for [
] i . Note that β i in this case is the non-NH 3 /
buffering power (i.e., the buffering power of all buffers other than NH 3 /
). If the NH 3 titration is performed in the absence of CO 2 /
, then the measured β i is
. If CO 2 /
is present, then the measured β i is
. An approach commonly used with
is to compute β from the pH i decrease caused by decreasing or withdrawing external NH 4 + . During such experiments, one usually blocks acid-extrusion mechanisms (e.g., by removing external Na + ) that would likely counteract the acidification. By reducing [
] o in a stepwise fashion ( Figure 52.9a ), one can compute β i as a function of pH i ( Figure 52.9b ), as has been done in mesangial cells and gastric parietal cells. An alternate approach one can use with
is to compute β i from the pH i increase that accompanies stepwise increases of external
.
Acid–base transporters can be divided into two groups: acid loaders and acid extruders. Acid loaders move H + into, or base equivalents (e.g., OH − or
) out of, cells and generally contribute to the recovery of pH i from acute intracellular alkali loads. In contrast, acid extruders move H + out of, or base equivalents into, cells and generally contribute to the recovery of pH i from acute intracellular acid loads. In the remainder of this section, we discuss major acid–base transporters in each of these two categories. For each transporter, we will examine at least the following three issues: (1) general function and molecular identity, (2) energetics and role in pH i regulation, and (3) effects of pharmacological agents. Please consult the following relevant chapters in this book on the molecular biology of Na + -coupled
transporters ( Chapter 53 ) and anion exchangers ( Chapter 54 ), as well as refs. on Na + H + exchangers.
As noted above, passive fluxes of H + , monovalent cationic acids (e.g.,
), and monovalent anionic bases (e.g.,
) are governed by the electrochemical gradient for the ion in question, and normally tend to acidify the cell. In principle, such passive transport processes could be exploited for protecting the cell against alkaline loads. Nevertheless, such passive processes generally do not appear to have a substantial impact on pH i . A notable exception discussed above is the GABA A -activated Cl − channel at the crayfish neuromuscular junction and in cells from turtle cerebellum and rat hippocampus. This Cl − channel can conduct
and thus mediate a substantial
efflux that lowers pH i . For the most part, however, major acid-loading transport pathways are
transporters—the most common ones being Cl −
exchange and Na + /
cotransport.
Found in a wide variety of animal-cell membranes, Cl −
exchangers normally couple the influx of Cl − to the efflux of
( Figure 52.10a–h ). Na + -independent Cl −
exchangers (to distinguish them from the Na + -dependent Cl −
exchangers discussed later) are thought to serve two major functions in non-epithelial cells: regulation of intracellular pH (pH i ), and regulation of intracellular [Cl − ]. The erythrocyte exchanger is known as the “band 3 protein” because of its position on SDS-polyacrylamide gels (for review, see ref. ) Band 3 has also been termed AE1 because it was the first “anion exchanger” to be cloned and sequenced ). As described in greater detail in Chapter 54 , the AE gene subfamily comprises at least three related genes AE1, AE2, and AE3. The AE gene subfamily is a branch of the bicarbonate-transporter (SLC4) family that includes the Na + -coupled
transporter subfamily (see Chapter 53 ). Several members of the SLC26 family can also mediate Cl −
exchange. However, unlike the AE family, which virtually always exhibits a Cl − :
stoichiometry of 1:1, the SLC26 family can exhibit stoichiometries of 1:2, 1:1, or 2:1 ; for review, see ref. .
Red cell AE1 mediates electroneutral homo- or hetero-exchange of monovalent anions with a 1:1 stoichiometry, exhibiting a substrate preference of Cl − ~
~NO 3 − >Br − >F − >I − >divalent oxyanions. Sulfate is cotransported with H + . Monovalent anion exchange can be reasonably well modeled by a ping-pong kinetic scheme, thought to reflect the alternation of an anion binding site between one side and the other of the membrane. Because the Cl − :
stoichiometry is 1:1, the transporter is electroneutral. Thus, the direction of net transport is determined by the sum of the chemical gradients for Cl − (i.e., [Cl − ] i /[Cl − ] o ) and
(i.e., [
] i /[
] o ). Because the inward Cl − gradient is generally greater than the inward
gradient in cells other than erythrocytes, the former dominates, driving
out of the cell. The transporter is easily reversed, however, by inverting the sum of the Cl − and
gradients. Indeed, a classic pH i assay for Cl −
exchange activity involves removing external Cl − that evokes a rapid pH i increase. In some cell types, the transporter can also be reversed by lowering [
] i , as would occur during severe intracellular acid loads. The exchanger operating in reverse then becomes an acid extruder, moving HCO 3 − into the cell and contributing to the pH i recovery from the acid load.
Some Cl −
exchangers have a very high affinity for
and transport can take place with low levels of
(~100 μM) in a nominally
-free solution. For example, removing external Cl − from the solution bathing rat osteoclasts can elicit a dramatic DIDS-sensitive pH i increase, even in the nominal absence of CO 2 /
. However, vigorously bubbling the solutions with 100% N 2 gas greatly reduces the pH i increase.
Sun et al. reported the presence of a Cl − OH − exchanger (or H + /Cl − cotransporter), which operates in parallel with a “conventional” Cl −
exchanger, in guinea-pig ventricular myocytes. The putative Cl − OH − exchanger is Cl − dependent; it is activated by low pH o and also by high pH i . The evidence that it is a Cl − OH − exchanger—rather than a Cl −
exchanger—is that it functions as an acid loader in the presence or absence of CO 2 /
, and is insensitive to DIDS. The activation of the exchanger by lowering pH o is unaffected by a CO 2 -free, 100% N 2 atmosphere. As attractive as it is to conclude that the transporter is a Cl − OH − exchanger, it is difficult to rule out the possibilities that: (1) ambient and metabolically produced CO 2 /
have not been totally eliminated, and (2) the transporter has an exceedingly high affinity for
. Putative Cl − OH − exchange activity also contributes to NaCl reabsorption in intestinal epithelial cells.
Based on experiments on rat-brain synaptosomes, Martínez-Zaguilán et al. have concluded that a H + /Cl − cotransporter is responsible for (1) the pH i increase elicited by removing extracellular Cl − , and (2) the faster pH i recovery from an acid load in the absence of extracellular Cl − . However, it is generally difficult to distinguish H + transport in one direction from OH − and/or
transport in the opposite direction.
The CLC gene family encodes proteins that are either Cl − channels or Cl − H + exchangers. While ClC1, ClC2, ClC-Ka, and ClC-Kb are plasma-membrane Cl − channels, ClC3 through ClC7 are thought to be vesicular Cl − H + exchangers. Although these ClCs are primarily expressed intracellularly, low levels of ClC4, ClC5, and ClC6 expression are observed at the plasma membrane, leading to Cl − H + exchanger activity at this location. The other ClCs in this group (i.e., ClC3 and ClC7) are also likely to be exchangers because each member of the group possesses a key glutamate residue that is necessary for such exchange.
By analogy to acid extruders, which are often stimulated by decreases in pH i , one might expect acid loaders such as Cl −
exchange to be stimulated by increases in pH i . Indeed, Cl −
exchange activity in many cells including Vero and L-cells, mesangial cells, and lymphocytes is low but measurable at pH i values as low as 6.5, increases to somewhat higher levels in the physiological pH i range, and then rises very steeply as pH i increases above ~7.6.
In more recent expression studies, the widely distributed AE2 also appears to exhibit greater activity at more alkaline pH i values. The pH sensor is located somewhere within the C-terminal transmembrane domain, and its apparent pK a is modulated by residues within the N-terminal cytoplasmic domain. Multiple histidines within the transmembrane domain contribute to AE2 activity, as well as pH i and pH o sensitivies. Furthermore, residues within the first putative re-entrant loop of AE2 also contribute to pH i and/or pH o sensitivities.
The non-erythroid Cl −
exchanger is blocked by stilbene derivatives such as DIDS and SITS, as well as the noncompetitive inhibitor niflumic acid. Some of the more potent AE inhibitors include oxonol dyes at the nanomolar level.
Become a Clinical Tree membership for Full access and enjoy Unlimited articles
If you are a member. Log in here