Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Oct 20.
Published in final edited form as: FEBS Lett. 2019 May 18;593(13):1528–1541. doi: 10.1002/1873-3468.13427

Mitochondrial Ca2+ concentrations in live cells: quantifications and discrepancies

Celia Fernandez-Sanz 1, Sergio De la Fuente 1,*, Shey-Shing Sheu 1,*
PMCID: PMC7573507  NIHMSID: NIHMS1637349  PMID: 31058316

Abstract

Intracellular Ca2+ signaling is a fundamental process that controls many cellular functions. During intracellular Ca2+ signaling mitochondria are capable of responding to cytosolic Ca2+ changes to regulate their own activities and, in some cell types, shape the spatiotemporal properties of the cytosolic Ca2+ signal. Numerous methods have been developed to specifically and quantitatively measure the mitochondrial free Ca2+ concentrations ([Ca2+]m), but there are still significant discrepancies in the absolute values of [Ca2+]m in live cells upon physiological stimulus. These discrepancies may derive from the diverse properties of the method used to measure [Ca2+]m, the calcium-free/bound ratio, and the intrinsic Ca2+ dynamics of each cell type in responding to various stimuli. Critical processes happening in the mitochondria, such as the ATP generation, ROS homeostasis, and mitochondrial permeability transition opening, depend directly on certain levels of the [Ca2+]m. Knowing those absolute [Ca2+]m values precisely is imperative for understanding Ca2+ signaling in cell. This review summarizes the reported calibrated [Ca2+]m values in many cell types and explains the discrepancies between these values. The gap areas for future research are also proposed.

Keywords: Mitochondrial Ca2+ Concentrations, Live Cells, Fluorescent Ca2+ Indicators, Genetically Encoded Ca2+ Indicators

INTRODUCTION

A general cellular response to extracellular stimulus is the rise in the cytosolic Ca2+ concentration ([Ca2+]c) leading to intracellular Ca2+ signaling which plays a key role in multiple cellular processes, including gene expression, differentiation, bioenergetics regulation, contraction, exocytosis, neurotransmission, and cell death [1]. Multiple pathways can lead to this [Ca2+]c increase, including the Ca2+ entry from the extracellular medium through the plasma membrane calcium channels and Ca2+ release from the intracellular Ca2+ stores like the endoplasmic/sarcoplasmic reticulum (ER/SR). The release of Ca2+ from the ER/SR can be mediated by depolarization-initiated mechanical coupling of L-type Ca channels with the type 1 ryanodine receptor (RyR1) or by Ca2+- or IP3-induced Ca2+ release, depending on the cell type and tissue. These pathways elevate the resting [Ca2+]c concentration from ~100 nM to a level that varies vastly from report to report (Table 1). Unlike other cellular second messengers, the Ca2+ can neither be created nor destroyed; Ca2+ can only be transported, released, or bound. Due to the importance of the [Ca2+]c increases in controlling cell functions, levels of [Ca2+]c need to be tightly regulated. During a Ca2+ transient, mitochondria can uptake Ca2+ from the cytosol to modulate the spatiotemporal profiles of [Ca2+]c mainly via the mitochondrial calcium uniporter complex (MCUC) or through other alternative pathways[2,3]. However, the increase in the mitochondrial Ca2+ free concentration ([Ca2+]m) is involved not only in controlling the cytosolic Ca2+ signal but also in regulating mitochondrial functions, such as energy metabolism[4]. Cellular activities are expected to change upon stimulation, a process that demands energy. Mitochondria utilize the increase in [Ca2+]m to enhance ATP production and subsequently the energy generation[5].

Table 1.

Reported mitochondrial calcium concentrations

PROBE Kd (μM) CELL TYPE STIMULUS [Ca2+]m reported References
Rhod-2 0.57 HeLa Histamine 3.5 μM [93]
Thapsigargin 900 nM
CCE 1.6 μM
Glomerulosa cells Extracellular Ca2+ 750nM Equilibrated with Cyto [94]
Fura-FF 5.50 RBL Ip3 10–20 μM [65,113]
Rat myocytes Electrical Stimulation 10 nM [96]
Rhod-5N 320 HeLa Histamine 20–80 μM [95]
Yc2 1.26 HeLa Histamine 3.2–10 μM [98]
Yc3 3.98 HeLa Histamine 49.5 μM
Yc4 104 HeLa Histamine 106 μM
Mitycam 0.2 Rabbit cardiomyocytes Electrical Stimulation 40 nM [99]
4mtD3cpv 0.47 Rat cardiomyocytes Electrical Stimulation 0.3–0.9 μM simulated [100]
Extracellular Ca2+ 0.3–2.4 μM simulated
AEQ variants 260–1000 Hela Histamine 20–150 μM [66,69,114116]
Adipocytes Bradykinin 2 μM [117]
TNBC Cells ATP 80 μM [118]
Human Fibroblast Histamine 45 μM [119]
Cromafin K+ 200–600 μM [101]
Cardiomyocytes Angiotensin 1 μM [120]
Pleura mesothelioma Bradykinin 50 μM [121]
SH-SY5Y neuroblastoma Bradykinin 50 μM [116]
N2a neuroblastoma Bradykinin 85 μM [122]
Smooth muscle cells 0 Na+ 3.5 μM [123]
ATP 2 μM
STH Huntington estriatial (virus) Extracellular Ca2+ 60–450 μM [124]
SOCE 25 μM

Because of its physiological and pathological relevance, mitochondrial Ca2+ dynamics have been studied intensively in the last decades. Methods to measure changes in [Ca2+]m have substantially evolved in recent years. These methods are essentially classified in two major groups: fluorescent dyes and genetically encoded Ca2+ indicators (GECIs). GECIs can be subdivided into fluorescent or bioluminescent probes. Ideally, all of these methods are suitable to accurately report [Ca2+]m values, independent of the cell type or the stimulus, however each has its own advantages and disadvantages, and different methods are preferred to measure [Ca2+]m depending on the cell type, stimulus, or even on the equipment used to perform the measurements. So, despite the advances in the number of tools available for quantifying [Ca2+]m, discrepancies persist in the absolute values that can be reached inside the mitochondrial matrix upon stimulation in live cells. It is possible that mitochondria from different cell types may reach significantly different [Ca2+]m, due to their own biochemical and physical properties (including localization and/or relation with other cellular organelles). However, under a given identical circumstance, different methods applied should report and confirm the same [Ca2+]m in the same cell type. In this review, we summarize the importance of knowing absolute values of [Ca2+]m and most used methods to achieve the measurements; and we compile most of the calibrated [Ca2+]m values reported in the literature. We also point out the discrepancies between these values and what may underlie these differences. Areas for future research are also proposed.

MITOCHONDRIAL Ca2+ AND BIOENERGETICS

Mitochondria are the main ATP producers through oxidative phosphorylation in eukaryotic cells. The reactions of the Krebs cycle and the utilization of the resulting reducing power by oxidative phosphorylation generates up to 32–36 ATP molecules, whereas glycolytic metabolism generates only 2. These reactions are regulated by [Ca2+]m in a mechanism designed to ensure that the ATP synthesis is coupled with the energy demand of the cell[6].

Due to the great capacity of mitochondria to uptake Ca2+, it was thought that these organelles were essential reservoirs of mobilizable intracellular Ca2+ ions[7]. Nevertheless, several independent studies from the 1980s[810] showed that total [Ca2+]m was low under basal conditions, but increases in [Ca2+]c in response to extrinsic agents were mirrored by increases of [Ca2+]m [11,12]. The higher [Ca2+]m in response to increases of [Ca2+]c would lead to the activation of oxidative metabolism and result in an increased supply of reducing equivalents to drive respiratory chain activity and ATP synthesis. Export of mitochondrial ATP in exchange for ADP will be anticipated to meet the higher ATP demand to fuel energy-requiring processes in the cytosol, such as ion pumping and contraction[13].

Mitochondrial enzymes modulated by [Ca2+]m

- Pyruvate dehydrogenase phosphatase (PDHP).

The pyruvate dehydrogenase (PDH) is a 50 MDa multi-enzyme complex that catalyzes the irreversible reaction of pyruvate, CoA, and NAD to obtain acetyl-CoA, NADH2, and CO2. The product, acetyl-CoA then enters to the citrate cycle or fatty acid synthesis. The regulation of the PDH complex activity is achieved via end-product inhibition and reversible inhibitory phosphorylation by a PDH kinase[14,15]. A Mg2+/Ca2+-dependent PDH phosphatase, in turn, dephosphorylates and reactivates PDH. It has been described that Ca2+ activates the PDH phosphatase in heart mitochondrial extracts with K0.5 of 0.7 μM[16,17]. Therefore, [Ca2+]m regulates the activity of the PDH complex, a rate-limiting step for pyruvate oxidation to acetyl CoA that in turn is reflected in the rate of ATP synthesis in mammalian tissues[6,18].

- 2-Oxoglutarate dehydrogenase (OGDH).

The OGDH is a sizeable multi-subunit complex, which similarly to the PDH complex, consists of multiple copies of three different subunits (E1/E2/E3). The complex catalyzes the multiple step reaction that transforms the 2-oxoglutarate to succinyl-CoA, Co2 and NADH[19]. From the many cofactors involved in the regulation of the OGDH complex, it has been found that Ca2+ and adenine nucleotides are important modulators that directly bind to the complex, most likely to the E1 subunit[20] with a described K0.5 of ∼1 μM in mitochondrial extracts and of ∼0.2 μM in permeabilized mitochondria. Additionally, it has been shown that high ADP/ATP ratios decrease not only the Km of the OGDH complex for the substrate but also the K0.5 for the Ca2+ [21]. It seems reasonable to think that under sudden high workload conditions, the high ADP/ATP mitochondrial ratios and [Ca2+]m increases over the same time scale[22,23] stimulate a reduced power supply for oxidative phosphorylation by increasing the activity of the OGDH.

- NAD+-isocitrate dehydrogenase (NAD-ICDH).

NAD-ICDH is a hetero-octamer of 2 × (2α,β,γ) subunits which catalyzed the oxidation of isocitrate to 2-ketoglutarate and CO2[24,25]. This enzyme, located exclusively in the mitochondria in mammalian cells, is a key regulatory enzyme in the TCA cycle. Calcium is one of the main cofactors of the NAD-ICDH. Importantly, the presence of ADP, isocitrate, and Mg2+ are required for Ca2+ to bind and regulate NAD-ICDH activity. The stoichiometry of Ca2+ binding to the enzyme is ∼1Ca2+ per tetramer (2α,β,γ)[11]. After binding to Ca2+ in the presence of ADP, the NAD-ICDH Km decreased ∼8-fold in mitochondrial extracts with a K0.5 for Ca2+ of 5.4 μM[21,26,27]. In the same direction, it has been observed that the K0.5 for Ca2+ decreased from 43 μM to 5 μM by increasing ADP/ATP ratios, in toluene-permeabilized mitochondria[21]. The higher K0.5 of the NAD-ICDH compared to the other intra-mitochondrial dehydrogenases (PDHP and OGDH) together with the thigh regulation of its activity depending on the ADP/ATP ratio might reflect here again how [Ca2+]m mediates the fine-tuning of the oxidative metabolism on these organelles in response to the different energy demands[12].

- F1–FO ATP synthase.

This molecular motor couples the ATP turnover with H+ translocation through the inner mitochondrial membrane. Therefore, the F1–FO ATP synthase utilizes the electrochemical proton gradient generated by the mitochondrial electron transport chain to synthesize ATP from ADP+Pi. In tissues with a high energy requirements like the myocardium, ATP synthesis has to match ATP consumption according to energy demands. It has been demonstrated by others that the regulation of the ATP production rate by the F1–FO ATP synthase might be independent from changes in mitochondrial membrane potential (Δψm) or rate of O2 consumption[28,29], being the enzyme the direct potential target of this regulation. The F1- F0 ATP synthase activity regulation by Ca2+ fluxes might explain the stabilizing of ATP production during workload that Ca2+-dehydrogenases regulation alone cannot explain. In this concern, Ca2+ has emerged as a key regulator of the ATP production. It has been reported that Ca2+ directly binds to F1–FO ATP synthase, specifically to the F1–FO ATP synthase β subunit[3032] with K0.5 for Ca2+ in the nM range[29]. Studies on isolated cardiac myocytes from rats have investigated [ATP] and [Ca2+] from both mitochondria and cytosol under different workload conditions. These measurements were performed by adenoviral infection on isolate cardiomyocytes with luciferase ([ATP]) and aequorin ([Ca2+]) specifically targeted to each cellular compartment[22]. Upon different workload conditions, no changes in the free [ATP] were observed in neither of the studied compartments. Interestingly, when cardiomyocytes were subjected to a sudden workload after a resting period, an initial drop in [ATP]m was observed before reaching equilibrium again. This drop in [ATP]m was accompanied with a simultaneous increase in [Ca2+]m indicating an ATP synthesis regulation mediated by [Ca2+]m variations in response to changes in myocardial workload[22].

- The mitochondrial permeability transition pore (mPTP).

Mitochondrial PTP is a non-specific channel that is known to allow the diffusion of molecules up to 1500 Da, but its molecular nature remains elusive [3345]. The mPTP has important physiological[46,47] and pathological roles[48,49]. Transient openings of the mPTP could participate in ROS generation[46,50,51] and Ca2+[52] handling. It has been postulated that the mPTP itself might also influence [Ca2+]m. Mathematical models suggest that under high intracellular calcium cycling during periods of stress, the mitochondrial calcium efflux pathway mediated by the mNCX may be unable to prevent calcium overload, suggesting possible different mechanisms of [Ca2+]m extrusion[52,53]. The observation of the mitochondrial calcium efflux inhibition in rat and mouse cardiomyocytes treated with mPTP blockers (cyclosporine A) or lacking cyclophilin D (CyP-D)[54,55] supported the possible role of mPTP on the mitochondrial calcium extrusion process. Furthermore, an elevated [Ca2+]m has been found to be related with a shift in substrate utilization from fatty acid oxidation to glycolysis in the working heart, suggesting that the mPTP might constitute a control point that links mitochondrial metabolism with myocardial workload through changes in the [Ca2+]m[53]. According to these observations, mPTP can modulate and sense changes in [Ca2+]m, acting as a mediator of cell energy demand adaptation or triggering cell death.

[Ca2+]m MEASUREMENT METHODS

As reported above, [Ca2+]m handling is essential in cell life. [Ca2+]m is not only essential for mitochondrial biogenetics and energy production but also crucial for buffering and shaping [Ca2+]c transients. Additionally, [Ca2+]m is directly involved in preventing or triggering cell death both by necrosis or apoptosis[56]. Therefore, it is important to accurately know the [Ca2+]m dynamics to fully understand all these physiological and pathological processes. Several methods are currently available to measure [Ca2+], from fluorescent dyes and proteins to bioluminescent sensors and radioisotopes of Ca2+. Those methods that are most frequently used to monitor mitochondrial Ca2+ dynamics in live cells and their pros and cons will be explained further in detail.

Fluorescent Dyes

A number of fluorescent dyes to measure intracellular [Ca2+] are currently available. All of them are based on the first fluorescent Ca2+ probe, Quin-2, developed back in the 1980s by Roger Tsien as a modification of the Ca2+ chelator EGTA[57]. A diverse group of Ca2+ fluorescent indicators has been developed since then, covering multiple combinations of spectral (single/double-wavelength or UV/visible excitation) and chemical (Ca2+ affinity, Kd) properties[58]. Some of these dyes are ratiometric indicators, like the Fura-2 and Indo-1 families. The Oregon Green family provides a resting signal while others like the Fluo family have no resting signal. Finally, the Rhod family are suitable to be used in the long wavelength range[59]. These dyes work with the principle of changing the fluorescence intensity according to changes in [Ca2+]. Most of them can be easily loaded into the cytosol. The non-invasive acetoxymethyl (AM) ester loading technique is the most popular to achieve this aim[60]. The acetoxymethyl ester group lends hydrophobic properties to the fluorescent dyes, allowing them to cross the cytosolic and mitochondrial membranes. To further promote mitochondrial accumulation, a positively charged molecule is then attached to the Ca2+ indicators (e.g., Rhod-2). The cationic nature of these dyes results in Δψm-driven uptake into the mitochondrial matrix. Once inside the matrix, mitochondrial esterases hydrolyze the AM groups, releasing the acids groups, retaining the probe in the mitochondrial matrix and conferring it the capability to respond to changes in the [Ca2+]m. The use of fluorescent dyes as Ca2+ sensors have many advantages. All of them are commercially available, and they are easy and fast to load in any type of cells, from established cell lines to primary cultures. The loading conditions must be optimized for each fluorescent dye, enabling high spatiotemporal resolution for imaging experiments. Most of them show a high Ca2+ affinity that goes from 0.1 μM to 1μM and a few show Ca2+ affinity at a low micromolar range (< 30 μM) and exceptionally at a micromolar range of > 30 μM. Despite of their versatility and easy handling, they also have strong disadvantages. Even with the AM technique, they are difficult to target and confine in the mitochondrial matrix. Incomplete AM ester hydrolysis, extracellular AM ester hydrolysis, and leakage are common problems related with these fluorescence dyes. The most concerning issue is that the dye, still localized in the cytosol, causes a strong signal contamination. Some strategies can be used to optimize the mitochondrial compartmentalization. The cold loading-warm incubation protocol[61] can help to reduce the artifactual cytosolic signal, and so can manganese treatments[62], cobalt treatments[63], and membrane permeabilization[64]. Despite the number of fluorescent dyes, Fura2-FF AM and Rhod-2 AM are the default selections to measure [Ca2+]m. The Fura2-FF is a ratiometric dye excited at 340/380 nm, with an emission peak at 510 nm. Thanks to its ratiometric condition, artifacts created during the acquisition can be avoided so that more reliable Ca2+ measurements can be obtained. According to the manufacturer, the Kd of Fura2-FF is 5μM, however, in vivo studies determined the Kd is approximately 4μM[65]. The Rhod-2 has a very high Ca2+ affinity (Kd = 570 nM), but some of his family members have lower affinity like Rhod-FF Kd = 19 μM or the extreme low Ca2+ affinity Rhod-5N with Kd = 350 μM. Previous studies have reported that Rhod family members are toxic for the cells when loaded at higher concentrations than 2 μM and suggest that Rhod-2 is unable to respond to two or more [Ca2+]m repetitive increases[66].

Genetically Encoded Calcium Indicators (GECIs)

Unlike the fluorescent dyes, GECIs are protein-like probes that can be easily targeted to different organelles by including a target sequence before the probe sequence. Another general advantage of the GECIs is that the sequence of their Ca2+-bound domain can be modified to reduce or adjust the probe affinity for Ca2+. This feature has allowed the generation of GECIs that cover a wide range of [Ca2+], from the low nM to high mM, making them suitable to perform [Ca2+] measurements in every cell organelle, including in the largest intracellular Ca2+ store, the ER/SR. The most significant disadvantage of GECIs is that they need to be expressed by the cell’s own machinery, meaning that the plasmid encoding for the probe needs to be delivered within the cell. Many cells lines do not present difficulties with transfection protocols, however others require different techniques, such as infection or electroporation. After the probe internalization, at least 24 – 48 h of probe expression is required prior to the experimental procedure. The waiting time for the expression of the probe is not a problem in cell lines, nevertheless it becomes critical when working with primary cells. Some of the primary cultures, like cardiomyocytes, are difficult to maintain without major structural changes for the time frame required for the probe’s expression. Alternative approaches exist to achieve expression of GECIs in primary cultures. Especially in rodent models, adenovirus or adeno-associated virus can be injected in the live animal to make the primary cell culture express the probe by the time the cell is isolated. Another disadvantage of GECIs is that, unlike with fluorescent dyes, the probe may be not present in every cell, a factor that can limit single cell imaging experiments. Two major types of GECIs have been developed, one type based on bioluminescent proteins and another type based on fluorescent proteins. Further individual characteristics and functional details will be provided in the following sections.

Bioluminescent based GECIs

The aequorin was the first organelle-targeted engineered Ca2+ indicator[67]. It was used for the first time during the 1960s and 1970s[68], and it is the most widely used bioluminescent Ca2+ probe. Aequorin emits light in a Ca2+-dependent reaction where the cofactor coelenterazine is oxidized, releasing CO2 and a single photon (470 nm) per aequorin molecule. Using a calibration curve, the amount of light emitted can be easily transferred to [Ca2+]. The native aequorin possesses 3 Ca2+ binding sites and a high Ca2+ affinity. This native version can only measure with reliability [Ca2+] between 0.1 μM and 5μM. Single point mutations were introduced in the first and second Ca2+ binding sites, reducing its Ca2+ affinity by ten times (disabling only the second binding site) or by 100 times (disabling both the first and second Ca2+ binding sites)[69]. The three aequorin variants (wtAEQ, mutAEQ, and 2mutAEQ) in combination with different types of coelenterazine allow the measurement of [Ca2+] from 0.1 μM to > 1 mM. Because of the bioluminescent nature of the aequorin, excitation illumination is not required; therefore, there are no autofluorescence or phototoxicity artefacts. However, the bioluminescent enzymatic reaction is irreversible in the experimental period, so the aequorin is “consumed”. This “consumption” was a problem when the aequorin was exposed to high [Ca2+], because it could lead to miscalculation of the real [Ca2+] values, however this problem was solved with the low Ca2+ affinity version of aequorin. The main disadvantage to aequorin is that the amount of light emitted is extremely low, and the experiments should be performed most of the times in a cell population. On the other hand, working with cell populations eliminates the cell-to-cell variability. Measurements of [Ca2+]m can be performed in individual cells expressing aequorin, and although the spatial resolution is very poor, the results are reliable due to its precise intracellular location[70]. The three aequorin variants mentioned above have been targeted to mitochondria, although the most frequently used is the mutAEQ in combination with the coelenterazine n. All the mitochondrial-targeted aequorins have been calibrated and report absolute [Ca2+]m values. Despite the fact that aequorin is the most commonly used bioluminescent Ca2+ probe, other chemiluminescent Ca2+ sensors are worth mentioning, including obelin[71], mitrocomin, photina[72], and clytin[73]. The recent advances in genetic engineering allowed the generation of other bioluminescent Ca2+ indicators, such as the RLuc8 based indicator BRAC[73], and the Split RLuc8-based Nano-Lantern[74]. Further details about these bioluminescent Ca2+ indicators were described in a review[75].

Fluorescent-based GECIs

Fluorescent-based GECIs emerged from three critical achievements: the discovery of the green fluorescent protein (GFP) [76]; the development of color variants suitable for fostering resonance energy transfer (FRET) between two fluorescent proteins[77]; and the finding that Ca2+ binding to the calmodulin (CaM) fused with the M13 peptide derived from the myosin light chain kinase[78]. The first developed FRET-based GECIs were FIP-CBSM and Cameleon. FIP-CBSM is characterized by the localization of the M13 peptide between BGFP and FGFP. Cameleon possesses yellow and cyan fluorescent proteins (FP) fused directly to the CaM and M13 peptide[79,80]. In this type of ratiometric GECI, Ca2+ binding causes a conformational change in the probe that alters the fluorescence intensity of both the donor and the acceptor and therefore the FRET efficiency.

Since their initial development, GECIs have been evolved in multiple directions to improve their detection efficacy. These improvements include the development of single-fluorescence Ca2+ sensors (Pericams, Camgaroos, GCaMPs, Cepias, and GECOs), the implementation of circulated permuted FP to enhance FRET-based indicators, and the remodeling the CaM-M13 interface (Dcpv family) or the replacing of the CaM as a Ca2+ sensor for troponin C (TN-indicators family) [8187]. Most of the GECIs use the CaM-M13 as Ca2+ binding site, but two fluorescent GECIs, the GFP-aequorin (GA) and the GFP and apo-aequorin (GAP), bind the Ca2+ to the aequorin [88,89]. The probes here work as a BRET system, where the photons emitted by aequorin are transferred to the GFP, as happens in jellyfish. Due to the massive list of fluorescence GECIs, detailed features of most of them have already been compiled in recent reviews [75,9092].

Fluorescent GECIs have some significant advantages over bioluminescent. The Ca2+ dynamics can be visualized with high resolution at the subcellular level due to their high fluorescence emission compared to the extremely low emission of photons by the bioluminescent probes. Another benefit over the bioluminescent probes is that the fluorescent reaction is reversible and does not require any cofactors. Despite recent improvements in fluorescent GECIs, the Ca2+ affinity in most of them is relatively high. Only a few display a Kd in the 1 μM to 10 μM range and even fewer > 10 μM. Many fluorescent GECIs from diverse families have been targeted to mitochondria and found to report reliable [Ca2+]m dynamics. Nevertheless, most of them have not been calibrated to absolute [Ca2+]m values.

REPORTED [Ca2+]m VALUES WITH DIFFERENT MEASUREMENT METHODS

Despite the abundance of tools to measure the [Ca2+]m, few researchers have calibrated the probes, so records reveal only qualitative changes in the [Ca2+]m. So while absolute [Ca2+]m values are critical for our understanding of Ca2+-mediated regulatory mechanisms of mitochondrial activities, ironically, they can barely be found in the published literature. The following paragraphs present the reported [Ca2+]m values found, the method used to measure [Ca2+]m, the cell type in which the experiments were performed, and the physiological stimulus applied to achieve the [Ca2+]m rise. In most of these experiments, intact cells were used. However, some studies in permeabilized cells are also mentioned. Although there are numerous reports on absolute values of [Ca2+]m in isolated mitochondria, these results are not discussed in the present review.

In 2001, Collings et al.[93] performed experiments in intact HeLa cells using 1μ M Rhod-2 as a probe. Their work reported a maximum [Ca2+]m of 3.5 μM upon histamine (100 μM) stimulation. The maximum [Ca2+]m reached was also measured upon ER Ca2+ leak (using the addition of thapsigargin to inhibit the SERCA pump) and during the capacitative Ca2+ entry. The values obtained were 0.4 μM and 1μM, respectively. These authors showed that even with prolonged stimulation (20 min) with histamine, [Ca2+]m barely reached 3μM. It is worth mentioning that the mitochondrial Rhod-2 fluorescent did not decrease significantly after 2 min following the maximum mitochondrial Ca2+ peak. The Rhod-2 signal was reduced to resting levels 25 min after the mitochondrial peak, pointing to the possibility that Rhod-2 cannot follow the mitochondrial Ca2+ dynamics properly. A mitochondrial calibration curve is also reported in the manuscript, showing a Kd of 1.3 μM for the Rhod-2. To obtain the minimum Rhod-2 fluorescence, 4 mM of EGTA was applied and a 10 μM Ca2+ pulse was used to reach the maximum Rhod-2 fluorescence.

In 2002, Pitter et al.[94] also used Rhod-2 (2 μM) to measure [Ca2+]m in permeabilized glomerulosa cells and in an INS-1/EK-3 cell line. The authors increased the [Ca2+]c progressively from 60 nM to 740 nM to determine mitochondrial Ca2+ uptake. The calibrated [Ca2+]m values were always similar to or even smaller than the [Ca2+]c added, despite a large electrochemical gradient favoring mitochondrial Ca2+ accumulation. This work used the calibration equation [Ca2+] = Kd (F-Fmin)/(Fmax-F)[58]. The basal fluorescence emission of the buffer with no Ca2+ added (no Ca2+ chelator present either) was consider as the minimum fluorescence. To obtain the maximum fluorescence, 500 μM Ca2+ was added to the buffer (in the presence of ionomycin and the uncoupler FCCP). A Kd value of 490 nM was reported for Rhod-2 and then used in the equation.

In 2012, De la Fuente et al.[95] used the extremely low Ca2+ affinity dye Rhod-5N to measure [Ca2+]m in intact HeLa cells. The stimulus used was histamine (100 μM) and histamine plus the MCUC activator Kaempferol. The [Ca2+]m values registered in individual cells (with histamine only) ranged from 10 to 40 μM, and the average of multiple cells was 30 μM. When the histamine was applied together with kaempferol, the average values were higher, reaching levels around 80 μM. Since the spatial resolution of the fluorescent dyes is high enough, [Ca2+]m values were also quantified in subcellular mitochondrial regions. The calibrated Rhod-5N signal showed values close to 120 μM, ~34 times higher than the value obtained by Collings et al., in those mitochondria surrounding the nucleus. Instead of using the Grynkiewicz calibration equation, De la Fuente et al. created an intracellular calibration curve, relating known [Ca2+] with normalized Rhod-5N fluorescence intensity. The calibration was done in the presence of FCCP, and the minimum fluorescent value was taken with 0.5 mM EGTA present and the maximum fluorescence with [Ca2+] 10 mM. The Kd reported in vivo for Rhod-5N in the described conditions was 470 μM.

In 2009, Andrienko et al.[96] used the high Ca2+ affinity dye Fura-2 (10 μM) and Rhod-2 (concentration not specified) to measure [Ca2+]m in permeabilized adult rat cardiomyocytes. Spontaneous SR Ca2+ release events occurring at regular frequency were used as mitochondrial Ca2+ uptake stimulus. According to the calibrated Fura-2 signals, [Ca2+]m increased to 110 nM after 8 waves and to 216 nM after 20 waves. The calibrated Rhod-2 signals were inferred from the obtained Fura-2 data. Therefore, the authors concluded that the largest [Ca2+]m rise during SR release was 10 nM, with the average rise in the 2 to 4 nM range. The Fura-2 calibration was done using Ca2+ free solution and Ca2+ 50 μM to 100 μM for minimum and maximum fluorescence, respectively; then the Grynkiewicz calibration equation was used (no Kd was reported).

In 2001, Csordas & Hajnoczky[65] utilized RBL-2H3 mucosal mast cells loaded with Fura-FF (5 μM) to measure mitochondrial Ca2+ uptake evoked by IP3-mediated ER Ca2+ release. The Kd obtained by a Ca2+ calibration kit (Calcium calibration kit2&3, molecular probes) was 4 μM, a similar value reported by others[97]. The imaging studies performed showed that in these conditions, [Ca2+]m reached 10–20 μM.

In the same year, Arnaudeau et al.[98] used three different variants of the fluorescent GECI Camaleon to measure [Ca2+]m in intact HeLa cells. The following Kd variants were reported: YC2mit 1.26 μM, YC3.1mit 3.98 μM, and YC4.1 104 μM. These Kd values were determined by applying increasing [Ca2+] from 1 nM to 10 mM (1 μM thapsigargin, 1 μM ionomycin, and 1 μM CCCP were present in the calibration buffer). The Hela cells were stimulated with histamine 50 μM, and the [Ca2+]m peak as wells as the % of saturated pixels were recorded. The study showed that [Ca2+]m increased as the Ca2+ affinity of the probe decreased. The Ca2+ values were 3.19 μM for YC2mit, 49.4 μM for YC3.1mit, and 106 μM for YC4.1mit. The saturated pixels were 24.8%, 17.8%, and 3.19% respectively, pointing to the fact that a low Kd affinity of the probe is critical to correctly and accurately measure [Ca2+]m values.

In 2012, Lu et al.[99] performed imaging experiments in intact adult rabbit cardiomyocytes using the fluorescent GECI Mitycam (inverse Pericam). At the end of every experiment, the Mitycam was calibrated in situ with low and high Ca2+ (concentrations not specified) to obtain the [Ca2+]m values. The results showed a gradient of Ca2+ within an individual mitochondrion with higher [Ca2+]m in the Z-lines (next to the SR Ca2+ releasing sites) than in M-lines. The reported [Ca2+]m were 37 nM in the Z-lines versus 26 nM in the M-lines. These values are higher but not significantly different from the ones obtained previously by the same group using Fura-2 and Rhod-2[96].

In 2017, Wust et al.[100] delivered the GECI 4mtD3cpv Camaleon (MitoCam) into adult rat cardiomyocytes by viral infection to estimate [Ca2+]m. Upon electrical stimulation (0.1–4 Hz), the [Ca2+]m values ranged from 300 nM to 800 nM while in presence of the Na+/Ca2+ exchanger inhibitor CGP-37157, the values vary from 580 nM to 1125 nM. The manuscript shows a MitoCam calibration where the Kd (470 nM) was applied to a modified Grynkiewicz equation[86].

Lastly, multiple publications in several cell types can be found where the bioluminescent GECI aequorin was used as a mitochondrial Ca2+ probe. The calibration of the different aequorins was performed as described by De la Fuente et al.[69]. Several Ca2+ concentrations were added in presence of thapsigargin, ionomycin, oligomycin and FCCP to obtain the calibration curve values. Ca2+ 10mM addition achieved the release of the total luminescence. The calibrated [Ca2+]m values from the aequorin and coelenterazine combinations differ significantly depending on the cell type used for the measurements. As such, the reported values of [Ca2+]m range from 1μM in cardiomyocytes to 600 μM in bovine chromaffin cells[101]. A list of reported [Ca2+]m using aequorins, including further details on the cell type, the stimulus, and the [Ca2+]m reached, can be found in Table 1.

POSSIBLE EXPLANATIONS FOR THE DISCREPANCIES IN [Ca2+]m REPORTED.

Despite many new tools and efforts to measure the [Ca2+]m accurately, there is still no consensus about the absolute values of [Ca2+] reached in the mitochondrial matrix during physiological stimulation, even in cases where the same cell types with the same stimulus were used. The discrepancies may result from a variety of factors. One of the most critical is the Kd of the probe for Ca2+. If the Kd is too low for the range of [Ca2+]m changes, then the dye will be saturated or close to the saturation; conversely, if the Kd is too high, then the changes may not be detected. This is because most of the probes follow a sigmoidal saturation curve, meaning that, when the Ca2+ levels start to deviate from Kd significantly, changes in the fluorescence intensity would be extremely low even when changes in [Ca2+] are still high. This phenomenon could lead to misinterpretation of the data since a non-saturated fluorescence signal does not necessarily translate to a correctly calibrated [Ca2+]. According to Arnaudeau et al.[98], higher [Ca2+]m values occurred when a lower Ca2+-affinity probe was used. Note that in this work none of the YC Camaleon signals were saturated at the [Ca2+]m peak. However, the variant with the lowest Kd resulted in measurement of the highest [Ca2+]m. The kinetics of the probes can also lead to different detecting signals, a factor that is especially critical for a rapid Ca2+ change, such as in beating cardiac muscle cells. If the probes are slower in picking up the real-time Ca2+ changes, then the signals will be distorted. Another source of discrepancy is the probe calibration performed by the different groups. Because calibration protocols are not standardized, every group carries out their own calibrations, leading to slight variations among them. One of the primary cause of discrepancies is that the Fmax and Fmin used for the Grynkiewicz equation are not always obtained under the same conditions. For the Fmin, some groups only omit the Ca2+ from the medium while others add the Ca2+ chelator EGTA in different concentrations to avoid any Ca2+ contamination. For the Fmax. a similar situation is faced. While some groups apply a [Ca2+] believed to saturate the probe, others go further and apply [Ca2+] in the 10 mM range to ensure the full saturation. Another possible explanation resides in the nature of the cells by themselves. The published data suggest a quite stable rise in the [Ca2+]c values (~ 1 μM) independently on the cell type, so the cytosolic Ca2+ is not a probable source of inconsistency. The high [Ca2+]c microdomains have not been as extensively studied as the global cytosolic Ca2+ signal[102]. Nevertheless, it does not appear to be the source of the variability either[103].In the other hand, mitochondria from diverse cell types may uptake Ca2+ in a different manner. The total amount of MCUC expressed in that particular cell type, or the proximity between mitochondria and the Ca2+ source may affect the mitochondrial Ca2+ uptake capabilities substantially[104]. Here is where the expression and the specific localization of the MCUC within the cell becomes critical, as does the concept of mitochondrial ‘contactology’ with other organelles, especially with the main intracellular Ca2+ reservoir, the ER/SR[105].

Another controversial point that may affect the reported free [Ca2+]m is the role of the mitochondrial Pi and its capacity to buffer and form Ca2+ phosphate precipitates[106]. It has been shown that the presence of phosphate facilitates the mitochondrial Ca2+ uptake leading to 10–30 times more Ca2+ accumulation[107]. However, this large Ca2+ mitochondrial load is simultaneously buffered by Pi having as a consequence mitochondrial morphological changes and Ca2+ precipitation[69]. While specific studies suggest that free [Ca2+]m higher than 2 μM sufficiently generates calcium phosphate precipitates[107], others claim that at least 1.5 mM Ca2+ is required to form the same precipitates[69]. These values align with the reported data on the solubility of the product of [HPO42-]•[Ca2+][108]. Additionally, the same study showed that the presence of 1 mM or 3 mM phosphate precludes [Ca2+]m to reach higher concentrations because of the above mentioned Ca2+ buffering capacity of the phosphate. A third study also showed that the presence of Pi reduced by almost 50% the [Ca2+]m reached (from an original value of 50 μM in the absence of Pi)[109]. The role of the Pi as a mitochondrial Ca2+ buffer as well as the still under debate calcium-free/bound ratio[107,110,111] creates an extremely complex scenario where accurate and precise measurement of [Ca2+]m is critical. Lastly, we can not forget that all Ca2+ sensors are Ca2+ buffers due to their Ca2+ binding capability. Therefore, for the sensors with a lower quantum yield, the higher loading concentrations could alter the original Ca2+ signals at some degree. However, it is hard to find studies reporting, discussing or making a comparison between the buffering capacities among the Ca2+ sensors[112].

CONCLUSIONS AND PERSPECTIVES

Mitochondrial Ca2+ uptake was first described almost 40 years ago. Alternations in [Ca2+]m regulate crucial processes such as enzyme activities, bioenergetics, and cell death. Despite the importance of the dynamic range of [Ca2+]m, the absolute values that can be reached upon stimulation are still not well known. The probes available for precisely measuring [Ca2+] inside the mitochondria have individual pros and cons. The ideal probe to measure [Ca2+]m requires some essential characteristics. It needs to be specifically and efficiently targeted to the mitochondrial matrix to avoid cytosolic signal contaminations. It should have fast enough kinetics, to be able to track both fast Ca2+ uptakes as well as fast Ca2+ release. It should have high quantum yields to avoid buffering. It should provide enough spatial resolution at both the cellular and subcellular levels. It needs to have the appropriate Ca2+ affinity to not to suffer saturation upon high [Ca2+]. Finally, it should be easy to deliver to the mitochondrial matrix, independently on the cell type. Further investigations are required to obtain this “perfect” mitochondrial Ca2+ sensor and use it to elucidate the [Ca2+]m reached upon physiological stimulation as well as the mitochondrial Ca2+ dynamics in the different cell types. As mentioned, mitochondria from different cell types may uptake Ca2+ in a quite variable range of concentrations. It is not possible to cover the full range of [Ca2+] changes (from <10 nM to >1mM) with only one “ideal” sensor. The decision of which one is better depends on all factors described above and the cell type where the [Ca2+]m will be measured. When its expression is feasible, we would recommend the use of fluorescent GECIs due to its specific organelle targeting. Among these, the ratiometric ones are preferred over the single-wavelength to ensure the interference of experimental artifacts. Accurate knowledge in these Ca2+ values will resolve current discrepancies in the amount of [Ca2+]m required to control essential cellular processes, including energy production and mPTP opening, and even to prevent the formation of calcium phosphate precipitates.

Acknowledgments

We thank Jennifer Wilson for English editing to this manuscript. This study was supported by: NIH/NHLBI (HL122124, HL093671, HL137266, HL137426, HL142864).

Footnotes

Disclosures

None.

REFERENCES

  • 1.Berridge MJ (2012) Calcium signalling remodelling and disease. Biochem Soc Trans 40, 297–309. [DOI] [PubMed] [Google Scholar]
  • 2.Campanella M, Pinton P & Rizzuto R (2004) Mitochondrial Ca2+ homeostasis in health and disease. Biol Res 37, 653–60. [DOI] [PubMed] [Google Scholar]
  • 3.De la Fuente S & Sheu S-S (2019) SR-mitochondria communication in adult cardiomyocytes: A close relationship where the Ca2+ has a lot to say. Arch. Biochem. Biophys 663, 259–268. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Giorgi C, Marchi S & Pinton P (2018) The machineries, regulation and cellular functions of mitochondrial calcium. Nat Rev Mol Cell Biol 19, 713–730. [DOI] [PubMed] [Google Scholar]
  • 5.Glancy B & Balaban RS (2012) Role of mitochondrial Ca2+ in the regulation of cellular energetics. Biochemistry 51, 2959–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Tarasov AI, Griffiths EJ & Rutter GA (2012) Regulation of ATP production by mitochondrial Ca2+. Cell Calcium 52, 28–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Lehninger AL, Carafoli E & Rossi CS (1967) Energy-linked ion movements in mitochondrial systems. Adv. Enzymol. Relat. Areas Mol. Biol 29, 259–320. [DOI] [PubMed] [Google Scholar]
  • 8.Denton RM & McCormack JG (1980) On the role of the calcium transport cycle in heart and other mammalian mitochondria. FEBS Lett. 119, 1–8. [DOI] [PubMed] [Google Scholar]
  • 9.Hansford RG & Castro F (1981) Effects of micromolar concentrations of free calcium ions on the reduction of heart mitochondrial NAD(P) by 2-oxoglutarate. Biochem. J 198, 525–533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Crompton M (1985) The Regulation of Mitochondrial Calcium Transport in Heart. Curr. Top. Membr. Transp 25, 231–276. [Google Scholar]
  • 11.Rutter GA & Denton RM (1989) The binding of Ca2+ ions to pig heart NAD+-isocitrate dehydrogenase and the 2-oxoglutarate dehydrogenase complex. Biochem. J 263, 453–462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Rutter GA (1990) Ca2(+)-binding to citrate cycle dehydrogenases. Int. J. Biochem 22, 1081–1088. [DOI] [PubMed] [Google Scholar]
  • 13.Jouaville LS, Pinton P, Bastianutto C, Rutter GA & Rizzuto R (1999) Regulation of mitochondrial ATP synthesis by calcium: evidence for a long-term metabolic priming. Proc. Natl. Acad. Sci. U. S. A 96, 13807–13812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Sale GJ & Randle PJ (1982) Occupancy of phosphorylation sites in pyruvate dehydrogenase phosphate complex in rat heart in vivo. Relation to proportion of inactive complex and rate of re-activation by phosphatase. Biochem. J. 206, 221–229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Fries M, Chauhan HJ, Domingo GJ, Jung H-I & Perham RN (2003) Site-directed mutagenesis of a loop at the active site of E1 (alpha2beta2) of the pyruvate dehydrogenase complex. A possible common sequence motif. Eur. J. Biochem 270, 861–870. [DOI] [PubMed] [Google Scholar]
  • 16.Denton RM, McCormack JG & Edgell NJ (1980) Role of calcium ions in the regulation of intramitochondrial metabolism. Effects of Na+, Mg2+ and ruthenium red on the Ca2+-stimulated oxidation of oxoglutarate and on pyruvate dehydrogenase activity in intact rat heart mitochondria. Biochem. J 190, 107–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Rutter GA, Midgley PJW & Denton RM (1989) Regulation of the pyruvate dehydrogenase complex by Ca2+ within toluene-permeabilized heart mitochondria. Biochim. Biophys. Acta BBA - Mol. Cell Res 1014, 263–270. [DOI] [PubMed] [Google Scholar]
  • 18.Randle PJ (1995) Metabolic fuel selection: general integration at the whole-body level. Proc. Nutr. Soc 54, 317–327. [DOI] [PubMed] [Google Scholar]
  • 19.MCLAIN AL, SZWEDA PA & SZWEDA LI (2011) α-Ketoglutarate dehydrogenase: A mitochondrial redox sensor. Free Radic. Res 45, 29–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Nichols BJ, Rigoulet M & Denton RM (1994) Comparison of the effects of Ca2+, adenine nucleotides and pH on the kinetic properties of mitochondrial NAD(+)-isocitrate dehydrogenase and oxoglutarate dehydrogenase from the yeast Saccharomyces cerevisiae and rat heart. Biochem. J 303 ( Pt 2), 461–465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Rutter GA & Denton RM (1988) Regulation of NAD+-linked isocitrate dehydrogenase and 2-oxoglutarate dehydrogenase by Ca2+ ions within toluene-permeabilized rat heart mitochondria. Interactions with regulation by adenine nucleotides and NADH/NAD+ ratios. Biochem. J 252, 181–189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Bell CJ, Bright NA, Rutter GA & Griffiths EJ (2006) ATP regulation in adult rat cardiomyocytes: time-resolved decoding of rapid mitochondrial calcium spiking imaged with targeted photoproteins. J. Biol. Chem 281, 28058–28067. [DOI] [PubMed] [Google Scholar]
  • 23.Griffiths EJ (2009) Mitochondrial calcium transport in the heart: physiological and pathological roles. J. Mol. Cell. Cardiol 46, 789–803. [DOI] [PubMed] [Google Scholar]
  • 24.Ramachandran N & Colman RF (1980) Chemical characterization of distinct subunits of pig heart DPN-specific isocitrate dehydrogenase. J. Biol. Chem 255, 8859–8864. [PubMed] [Google Scholar]
  • 25.Martínez-Rivas JM & Vega JM (1998) Purification and Characterization of NAD-Isocitrate Dehydrogenase from Chlamydomonas reinhardtii. Plant Physiol. 118, 249–255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Denton RM, Richards DA & Chin JG (1978) Calcium ions and the regulation of NAD+-linked isocitrate dehydrogenase from the mitochondria of rat heart and other tissues. Biochem. J 176, 899–906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Gabriel JL & Plaut GW (1984) Inhibition of bovine heart NAD-specific isocitrate dehydrogenase by reduced pyridine nucleotides: modulation of inhibition by ADP, NAD+, Ca2+, citrate, and isocitrate. Biochemistry 23, 2773–2778. [DOI] [PubMed] [Google Scholar]
  • 28.Mildaziene V, Baniene R, Nauciene Z, Marcinkeviciute A, Morkuniene R, Borutaite V, Kholodenko B & Brown GC (1996) Ca2+ stimulates both the respiratory and phosphorylation subsystems in rat heart mitochondria. Biochem. J 320 ( Pt 1), 329–334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Territo PR, Mootha VK, French SA & Balaban RS (2000) Ca(2+) activation of heart mitochondrial oxidative phosphorylation: role of the F(0)/F(1)-ATPase. Am. J. Physiol. Cell Physiol 278, C423–435. [DOI] [PubMed] [Google Scholar]
  • 30.Hubbard MJ & McHugh NJ (1996) Mitochondrial ATP synthase F1-β-subunit is a calcium-binding protein. FEBS Lett. 391, 323–329. [DOI] [PubMed] [Google Scholar]
  • 31.Giorgio V, Burchell V, Schiavone M, Bassot C, Minervini G, Petronilli V, Argenton F, Forte M, Tosatto S, Lippe G & Bernardi P (2017) Ca2+ binding to F‐ATP synthase β subunit triggers the mitochondrial permeability transition. EMBO Rep. 18, 1065–1076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Williams GSB, Boyman L & Lederer WJ (2015) Mitochondrial calcium and the regulation of metabolism in the heart. J. Mol. Cell. Cardiol 78, 35–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Zoratti M & Szabó I (1994) Electrophysiology of the inner mitochondrial membrane. J. Bioenerg. Biomembr. 26, 543–553. [DOI] [PubMed] [Google Scholar]
  • 34.Crompton M, Virji S & Ward JM (1998) Cyclophilin-D binds strongly to complexes of the voltage-dependent anion channel and the adenine nucleotide translocase to form the permeability transition pore. Eur. J. Biochem 258, 729–735. [DOI] [PubMed] [Google Scholar]
  • 35.Baines CP (2010) The cardiac mitochondrion: nexus of stress. Annu. Rev. Physiol. 72, 61–80. [DOI] [PubMed] [Google Scholar]
  • 36.Kokoszka JE, Waymire KG, Levy SE, Sligh JE, Cai J, Jones DP, MacGregor GR & Wallace DC (2004) The ADP/ATP translocator is not essential for the mitochondrial permeability transition pore. Nature 427, 461–465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Santos AC, Uyemura SA, Lopes JL, Bazon JN, Mingatto FE & Curti C (1998) Effect of naturally occurring flavonoids on lipid peroxidation and membrane permeability transition in mitochondria. Free Radic. Biol. Med 24, 1455–1461. [DOI] [PubMed] [Google Scholar]
  • 38.Wang X-J, Wang Z-B & Xu J-X (2005) Effect of salvianic acid A on lipid peroxidation and membrane permeability in mitochondria. J. Ethnopharmacol 97, 441–445. [DOI] [PubMed] [Google Scholar]
  • 39.Kuwana T & Newmeyer DD (2003) Bcl-2-family proteins and the role of mitochondria in apoptosis. Curr. Opin. Cell Biol 15, 691–699. [DOI] [PubMed] [Google Scholar]
  • 40.Basañez G, Soane L & Hardwick JM (2012) A New View of the Lethal Apoptotic Pore. PLoS Biol. 10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.He L & Lemasters JJ (2002) Regulated and unregulated mitochondrial permeability transition pores: a new paradigm of pore structure and function? FEBS Lett. 512, 1–7. [DOI] [PubMed] [Google Scholar]
  • 42.Giorgio V, von Stockum S, Antoniel M, Fabbro A, Fogolari F, Forte M, Glick GD, Petronilli V, Zoratti M, Szabó I, Lippe G & Bernardi P (2013) Dimers of mitochondrial ATP synthase form the permeability transition pore. Proc. Natl. Acad. Sci. U. S. A 110, 5887–5892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Alavian KN, Beutner G, Lazrove E, Sacchetti S, Park H-A, Licznerski P, Li H, Nabili P, Hockensmith K, Graham M, Porter GA & Jonas EA (2014) An uncoupling channel within the c-subunit ring of the F1FO ATP synthase is the mitochondrial permeability transition pore. Proc. Natl. Acad. Sci. U. S. A 111, 10580–10585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Bernardi P & Di Lisa F (2015) The mitochondrial permeability transition pore: molecular nature and role as a target in cardioprotection. J. Mol. Cell. Cardiol 78, 100–106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Karch J, Bround MJ, Khalil H, Sargent MA, Latchman N, Terada N, Peixoto PM & Molkentin JD (2018) Inhibition of mitochondrial permeability transition by deletion of the ANT family and CypD. bioRxiv, 506964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Gong G, Liu X, Zhang H, Sheu S-S & Wang W (2015) Mitochondrial flash as a novel biomarker of mitochondrial respiration in the heart. Am. J. Physiol. - Heart Circ. Physiol 309, H1166–H1177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Hom JR, Quintanilla RA, Hoffman DL, de Mesy Bentley KL, Molkentin JD, Sheu S-S & Porter GA (2011) The permeability transition pore controls cardiac mitochondrial maturation and myocyte differentiation. Dev. Cell 21, 469–478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Paradies G, Ruggiero FM, Petrosillo G & Quagliariello E (1993) Age-dependent decrease in the cytochrome c oxidase activity and changes in phospholipids in rat-heart mitochondria. Arch. Gerontol. Geriatr 16, 263–272. [DOI] [PubMed] [Google Scholar]
  • 49.Ruiz-Meana M, Abellán A, Miró-Casas E & Garcia-Dorado D (2007) Opening of mitochondrial permeability transition pore induces hypercontracture in Ca2+ overloaded cardiac myocytes. Basic Res. Cardiol 102, 542–552. [DOI] [PubMed] [Google Scholar]
  • 50.Zorov DB, Juhaszova M & Sollott SJ (2014) Mitochondrial Reactive Oxygen Species (ROS) and ROS-Induced ROS Release. Physiol. Rev 94, 909–950. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Brookes PS, Yoon Y, Robotham JL, Anders MW & Sheu S-S (2004) Calcium, ATP, and ROS: a mitochondrial love-hate triangle. Am. J. Physiol. Cell Physiol 287, C817–833. [DOI] [PubMed] [Google Scholar]
  • 52.Bernardi P & von Stockum S (2012) The permeability transition pore as a Ca(2+) release channel: new answers to an old question. Cell Calcium 52, 22–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Elrod JW & Molkentin JD (2013) Physiologic functions of cyclophilin D and the mitochondrial permeability transition pore. Circ. J. Off. J. Jpn. Circ. Soc 77, 1111–1122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Altschuld RA, Hohl CM, Castillo LC, Garleb AA, Starling RC & Brierley GP (1992) Cyclosporin inhibits mitochondrial calcium efflux in isolated adult rat ventricular cardiomyocytes. Am. J. Physiol 262, H1699–1704. [DOI] [PubMed] [Google Scholar]
  • 55.Elrod JW, Wong R, Mishra S, Vagnozzi RJ, Sakthievel B, Goonasekera SA, Karch J, Gabel S, Farber J, Force T, Brown JH, Murphy E & Molkentin JD (2010) Cyclophilin D controls mitochondrial pore-dependent Ca(2+) exchange, metabolic flexibility, and propensity for heart failure in mice. J. Clin. Invest 120, 3680–3687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Orrenius S, Gogvadze V & Zhivotovsky B (2015) Calcium and mitochondria in the regulation of cell death. Biochem. Biophys. Res. Commun 460, 72–81. [DOI] [PubMed] [Google Scholar]
  • 57.Tsien RY, Pozzan T & Rink TJ (1984) Calcium activities and fluxes inside small intact cells as measured with intracellularly trapped chelators. Adv Cycl. Nucleotide Protein Phosphorylation Res 17, 535–41. [PubMed] [Google Scholar]
  • 58.Grynkiewicz G, Poenie M & Tsien RY (1985) A new generation of Ca2+ indicators with greatly improved fluorescence properties. J Biol Chem 260, 3440–50. [PubMed] [Google Scholar]
  • 59.Minta A, Kao JP & Tsien RY (1989) Fluorescent indicators for cytosolic calcium based on rhodamine and fluorescein chromophores. J Biol Chem 264, 8171–8. [PubMed] [Google Scholar]
  • 60.Tsien RY (1981) A non-disruptive technique for loading calcium buffers and indicators into cells. Nature 290, 527–8. [DOI] [PubMed] [Google Scholar]
  • 61.Trollinger DR, Cascio WE & Lemasters JJ (2000) Mitochondrial calcium transients in adult rabbit cardiac myocytes: inhibition by ruthenium red and artifacts caused by lysosomal loading of Ca(2+)-indicating fluorophores. Biophys J 79, 39–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Miyata H, Silverman HS, Sollott SJ, Lakatta EG, Stern MD & Hansford RG (1991) Measurement of mitochondrial free Ca2+ concentration in living single rat cardiac myocytes. Am. J. Physiol 261, H1123–1134. [DOI] [PubMed] [Google Scholar]
  • 63.Davidson SM, Yellon D & Duchen MR (2007) Assessing mitochondrial potential, calcium, and redox state in isolated mammalian cells using confocal microscopy. Methods Mol. Biol. Clifton NJ 372, 421–430. [DOI] [PubMed] [Google Scholar]
  • 64.Sedova M, Dedkova EN & Blatter LA (2006) Integration of rapid cytosolic Ca2+ signals by mitochondria in cat ventricular myocytes. Am. J. Physiol. Cell Physiol 291, C840–850. [DOI] [PubMed] [Google Scholar]
  • 65.Csordas G & Hajnoczky G (2001) Sorting of calcium signals at the junctions of endoplasmic reticulum and mitochondria. Cell Calcium 29, 249–62. [DOI] [PubMed] [Google Scholar]
  • 66.Fonteriz RI, de la Fuente S, Moreno A, Lobaton CD, Montero M & Alvarez J (2010) Monitoring mitochondrial [Ca(2+)] dynamics with rhod-2, ratiometric pericam and aequorin. Cell Calcium 48, 61–9. [DOI] [PubMed] [Google Scholar]
  • 67.Rizzuto R, Simpson AW, Brini M & Pozzan T (1992) Rapid changes of mitochondrial Ca2+ revealed by specifically targeted recombinant aequorin. Nature 358, 325–7. [DOI] [PubMed] [Google Scholar]
  • 68.Shimomura O, Johnson FH & Saiga Y (1963) Microdetermination of Calcium by Aequorin Luminescence. Science 140, 1339–40. [DOI] [PubMed] [Google Scholar]
  • 69.de la Fuente S, Fonteriz RI, de la Cruz PJ, Montero M & Alvarez J (2012) Mitochondrial free [Ca(2+)] dynamics measured with a novel low-Ca(2+) affinity aequorin probe. Biochem J 445, 371–6. [DOI] [PubMed] [Google Scholar]
  • 70.Nunez L, Senovilla L, Sanz-Blasco S, Chamero P, Alonso MT, Villalobos C & Garcia-Sancho J (2007) Bioluminescence imaging of mitochondrial Ca2+ dynamics in soma and neurites of individual adult mouse sympathetic neurons. J Physiol 580, 385–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Campbell AK (1974) Extraction, partial purification and properties of obelin, the calcium-activated luminescent protein from the hydroid Obelia geniculata. Biochem J 143, 411–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Bovolenta S, Foti M, Lohmer S & Corazza S (2007) Development of a Ca(2+)-activated photoprotein, Photina, and its application to high-throughput screening. J Biomol Screen 12, 694–704. [DOI] [PubMed] [Google Scholar]
  • 73.Inouye S & Tsuji FI (1993) Cloning and sequence analysis of cDNA for the Ca(2+)-activated photoprotein, clytin. FEBS Lett 315, 343–6. [DOI] [PubMed] [Google Scholar]
  • 74.Saito K, Chang YF, Horikawa K, Hatsugai N, Higuchi Y, Hashida M, Yoshida Y, Matsuda T, Arai Y & Nagai T (2012) Luminescent proteins for high-speed single-cell and whole-body imaging. Nat Commun 3, 1262. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Perez Koldenkova V & Nagai T (2013) Genetically encoded Ca(2+) indicators: properties and evaluation. Biochim Biophys Acta 1833, 1787–97. [DOI] [PubMed] [Google Scholar]
  • 76.Shimomura O, Johnson FH & Saiga Y (1962) Extraction, purification and properties of aequorin, a bioluminescent protein from the luminous hydromedusan, Aequorea. J Cell Comp Physiol 59, 223–39. [DOI] [PubMed] [Google Scholar]
  • 77.Heim R & Tsien RY (1996) Engineering green fluorescent protein for improved brightness, longer wavelengths and fluorescence resonance energy transfer. Curr Biol 6, 178–82. [DOI] [PubMed] [Google Scholar]
  • 78.Porumb T, Yau P, Harvey TS & Ikura M (1994) A calmodulin-target peptide hybrid molecule with unique calcium-binding properties. Protein Eng 7, 109–15. [DOI] [PubMed] [Google Scholar]
  • 79.Romoser VA, Hinkle PM & Persechini A (1997) Detection in living cells of Ca2+-dependent changes in the fluorescence emission of an indicator composed of two green fluorescent protein variants linked by a calmodulin-binding sequence. A new class of fluorescent indicators. J Biol Chem 272, 13270–4. [DOI] [PubMed] [Google Scholar]
  • 80.Miyawaki A, Llopis J, Heim R, McCaffery JM, Adams JA, Ikura M & Tsien RY (1997) Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin. Nature 388, 882–7. [DOI] [PubMed] [Google Scholar]
  • 81.Nagai T, Sawano A, Park ES & Miyawaki A (2001) Circularly permuted green fluorescent proteins engineered to sense Ca2+. Proc. Natl. Acad. Sci. U. S. A. 98, 3197–3202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Baird GS, Zacharias DA & Tsien RY (1999) Circular permutation and receptor insertion within green fluorescent proteins. Proc Natl Acad Sci U A 96, 11241–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Nakai J, Ohkura M & Imoto K (2001) A high signal-to-noise Ca(2+) probe composed of a single green fluorescent protein. Nat Biotechnol 19, 137–41. [DOI] [PubMed] [Google Scholar]
  • 84.Suzuki J, Kanemaru K, Ishii K, Ohkura M, Okubo Y & Iino M (2014) Imaging intraorganellar Ca2+ at subcellular resolution using CEPIA. Nat Commun 5, 4153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Zhao Y, Araki S, Wu J, Teramoto T, Chang YF, Nakano M, Abdelfattah AS, Fujiwara M, Ishihara T, Nagai T & Campbell RE (2011) An expanded palette of genetically encoded Ca(2)(+) indicators. Science 333, 1888–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Palmer AE, Giacomello M, Kortemme T, Hires SA, Lev-Ram V, Baker D & Tsien RY (2006) Ca2+ indicators based on computationally redesigned calmodulin-peptide pairs. Chem Biol 13, 521–30. [DOI] [PubMed] [Google Scholar]
  • 87.Reiff DF, Ihring A, Guerrero G, Isacoff EY, Joesch M, Nakai J & Borst A (2005) In vivo performance of genetically encoded indicators of neural activity in flies. J Neurosci 25, 4766–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Baubet V, Le Mouellic H, Campbell AK, Lucas-Meunier E, Fossier P & Brulet P (2000) Chimeric green fluorescent protein-aequorin as bioluminescent Ca2+ reporters at the single-cell level. Proc Natl Acad Sci U A 97, 7260–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Rodriguez-Garcia A, Rojo-Ruiz J, Navas-Navarro P, Aulestia FJ, Gallego-Sandin S, Garcia-Sancho J & Alonso MT (2014) GAP, an aequorin-based fluorescent indicator for imaging Ca2+ in organelles. Proc Natl Acad Sci U A 111, 2584–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Suzuki J, Kanemaru K & Iino M (2016) Genetically Encoded Fluorescent Indicators for Organellar Calcium Imaging. Biophys J 111, 1119–1131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Alonso MT, Rodriguez-Prados M, Navas-Navarro P, Rojo-Ruiz J & Garcia-Sancho J (2017) Using aequorin probes to measure Ca(2+) in intracellular organelles. Cell Calcium 64, 3–11. [DOI] [PubMed] [Google Scholar]
  • 92.Granatiero V, Patron M, Tosatto A, Merli G & Rizzuto R (2014) The use of aequorin and its variants for Ca2+ measurements. Cold Spring Harb Protoc 2014, 9–16. [DOI] [PubMed] [Google Scholar]
  • 93.Collins TJ, Lipp P, Berridge MJ & Bootman MD (2001) Mitochondrial Ca(2+) uptake depends on the spatial and temporal profile of cytosolic Ca(2+) signals. J Biol Chem 276, 26411–20. [DOI] [PubMed] [Google Scholar]
  • 94.Pitter JG, Maechler P, Wollheim CB & Spat A (2002) Mitochondria respond to Ca2+ already in the submicromolar range: correlation with redox state. Cell Calcium 31, 97–104. [DOI] [PubMed] [Google Scholar]
  • 95.de la Fuente S, Fonteriz RI, Montero M & Alvarez J (2012) Dynamics of mitochondrial [Ca(2+)] measured with the low-Ca(2+)-affinity dye rhod-5N. Cell Calcium 51, 65–71. [DOI] [PubMed] [Google Scholar]
  • 96.Andrienko TN, Picht E & Bers DM (2009) Mitochondrial free calcium regulation during sarcoplasmic reticulum calcium release in rat cardiac myocytes. J Mol Cell Cardiol 46, 1027–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Xu-Friedman MA & Regehr WG (1999) Presynaptic strontium dynamics and synaptic transmission. Biophys J 76, 2029–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Arnaudeau S, Kelley WL, Walsh JV Jr & Demaurex N (2001) Mitochondria recycle Ca(2+) to the endoplasmic reticulum and prevent the depletion of neighboring endoplasmic reticulum regions. J Biol Chem 276, 29430–9. [DOI] [PubMed] [Google Scholar]
  • 99.Lu X, Ginsburg KS, Kettlewell S, Bossuyt J, Smith GL & Bers DM (2013) Measuring local gradients of intramitochondrial [Ca(2+)] in cardiac myocytes during sarcoplasmic reticulum Ca(2+) release. Circ. Res. 112, 424–431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Wust RC, Helmes M, Martin JL, van der Wardt TJ, Musters RJ, van der Velden J & Stienen GJ (2017) Rapid frequency-dependent changes in free mitochondrial calcium concentration in rat cardiac myocytes. J Physiol 595, 2001–2019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Montero M, Alonso MT, Carnicero E, Cuchillo-Ibanez I, Albillos A, Garcia AG, Garcia-Sancho J & Alvarez J (2000) Chromaffin-cell stimulation triggers fast millimolar mitochondrial Ca2+ transients that modulate secretion. Nat Cell Biol 2, 57–61. [DOI] [PubMed] [Google Scholar]
  • 102.Csordás G, Várnai P, Golenár T, Roy S, Purkins G, Schneider TG, Balla T & Hajnóczky G (2010) Imaging Interorganelle Contacts and Local Calcium Dynamics at the ER-Mitochondrial Interface. Mol. Cell 39, 121–132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Rizzuto R & Pozzan T (2006) Microdomains of intracellular Ca2+: molecular determinants and functional consequences. Physiol. Rev. 86, 369–408. [DOI] [PubMed] [Google Scholar]
  • 104.Fieni F, Lee SB, Jan YN & Kirichok Y (2012) Activity of the mitochondrial calcium uniporter varies greatly between tissues. Nat Commun 3, 1317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Csordas G, Weaver D & Hajnoczky G (2018) Endoplasmic Reticulum-Mitochondrial Contactology: Structure and Signaling Functions. Trends Cell Biol 28, 523–540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Nicholls DG (2005) Mitochondria and calcium signaling. Cell Calcium 38, 311–7. [DOI] [PubMed] [Google Scholar]
  • 107.Chalmers S & Nicholls DG (2003) The relationship between free and total calcium concentrations in the matrix of liver and brain mitochondria. J Biol Chem 278, 19062–70. [DOI] [PubMed] [Google Scholar]
  • 108.Fryer MW, Owen VJ, Lamb GD & Stephenson DG (1995) Effects of creatine phosphate and P(i) on Ca2+ movements and tension development in rat skinned skeletal muscle fibres. J Physiol 482 ( Pt 1), 123–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Wei AC, Liu T & O’Rourke B (2015) Dual Effect of Phosphate Transport on Mitochondrial Ca2+ Dynamics. J Biol Chem 290, 16088–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Babcock DF, Herrington J, Goodwin PC, Park YB & Hille B (1997) Mitochondrial participation in the intracellular Ca2+ network. J Cell Biol 136, 833–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.de la Fuente S, Montenegro P, Fonteriz RI, Moreno A, Lobatón CD, Montero M & Alvarez J (2010) The dynamics of mitochondrial Ca2+ fluxes. Biochim. Biophys. Acta 1797, 1727–1735. [DOI] [PubMed] [Google Scholar]
  • 112.McMahon SM & Jackson MB (2018) An Inconvenient Truth: Calcium Sensors Are Calcium Buffers. Trends Neurosci. 41, 880–884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Csordas G, Thomas AP & Hajnoczky G (1999) Quasi-synaptic calcium signal transmission between endoplasmic reticulum and mitochondria. Embo J 18, 96–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Cieri D, Vicario M, Vallese F, D’Orsi B, Berto P, Grinzato A, Catoni C, De Stefani D, Rizzuto R, Brini M & Cali T (2018) Tau localises within mitochondrial sub-compartments and its caspase cleavage affects ER-mitochondria interactions and cellular Ca(2+) handling. Biochim Biophys Acta Mol Basis Dis 1864, 3247–3256. [DOI] [PubMed] [Google Scholar]
  • 115.Patron M, Checchetto V, Raffaello A, Teardo E, Vecellio Reane D, Mantoan M, Granatiero V, Szabo I, De Stefani D & Rizzuto R (2014) MICU1 and MICU2 finely tune the mitochondrial Ca2+ uniporter by exerting opposite effects on MCU activity. Mol Cell 53, 726–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Cali T, Ottolini D, Negro A & Brini M (2012) alpha-Synuclein controls mitochondrial calcium homeostasis by enhancing endoplasmic reticulum-mitochondria interactions. J Biol Chem 287, 17914–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Wright LE, Vecellio Reane D, Milan G, Terrin A, Di Bello G, Belligoli A, Sanna M, Foletto M, Favaretto F, Raffaello A, Mammucari C, Nitti D, Vettor R & Rizzuto R (2017) Increased mitochondrial calcium uniporter in adipocytes underlies mitochondrial alterations associated with insulin resistance. Am J Physiol Endocrinol Metab 313, E641–e650. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Tosatto A, Sommaggio R, Kummerow C, Bentham RB, Blacker TS, Berecz T, Duchen MR, Rosato A, Bogeski I, Szabadkai G, Rizzuto R & Mammucari C (2016) The mitochondrial calcium uniporter regulates breast cancer progression via HIF-1alpha. EMBO Mol Med 8, 569–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Granatiero V, Giorgio V, Cali T, Patron M, Brini M, Bernardi P, Tiranti V, Zeviani M, Pallafacchina G, De Stefani D & Rizzuto R (2016) Reduced mitochondrial Ca(2+) transients stimulate autophagy in human fibroblasts carrying the 13514A>G mutation of the ND5 subunit of NADH dehydrogenase. Cell Death Differ 23, 231–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Robert V, Gurlini P, Tosello V, Nagai T, Miyawaki A, Di Lisa F & Pozzan T (2001) Beat-to-beat oscillations of mitochondrial [Ca2+] in cardiac cells. Embo J 20, 4998–5007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Patergnani S, Giorgi C, Maniero S, Missiroli S, Maniscalco P, Bononi I, Martini F, Cavallesco G, Tognon M & Pinton P (2015) The endoplasmic reticulum mitochondrial calcium cross talk is downregulated in malignant pleural mesothelioma cells and plays a critical role in apoptosis inhibition. Oncotarget 6, 23427–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Coussee E, De Smet P, Bogaert E, Elens I, Van Damme P, Willems P, Koopman W, Van Den Bosch L & Callewaert G (2011) G37R SOD1 mutant alters mitochondrial complex I activity, Ca(2+) uptake and ATP production. Cell Calcium 49, 217–25. [DOI] [PubMed] [Google Scholar]
  • 123.Poburko D, Potter K, van Breemen E, Fameli N, Liao CH, Basset O, Ruegg UT & van Breemen C (2006) Mitochondria buffer NCX-mediated Ca2+-entry and limit its diffusion into vascular smooth muscle cells. Cell Calcium 40, 359–71. [DOI] [PubMed] [Google Scholar]
  • 124.Lim D, Bertoli A, Sorgato MC & Moccia F (2016) Generation and usage of aequorin lentiviral vectors for Ca(2+) measurement in sub-cellular compartments of hard-to-transfect cells. Cell Calcium 59, 228–39. [DOI] [PubMed] [Google Scholar]

RESOURCES