Next Article in Journal
Mitophagy in Cancer: A Tale of Adaptation
Next Article in Special Issue
The Zinc Transporter Zip7 Is Downregulated in Skeletal Muscle of Insulin-Resistant Cells and in Mice Fed a High-Fat Diet
Previous Article in Journal
Anti-Inflammatory Effects by Pharmacological Inhibition or Knockdown of Fatty Acid Amide Hydrolase in BV2 Microglial Cells
Previous Article in Special Issue
The Copper(II)-Assisted Connection between NGF and BDNF by Means of Nerve Growth Factor-Mimicking Short Peptides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Live-Cell Imaging of Physiologically Relevant Metal Ions Using Genetically Encoded FRET-Based Probes

by
Helmut Bischof
1,*,
Sandra Burgstaller
1,
Markus Waldeck-Weiermair
1,
Thomas Rauter
1,
Maximilian Schinagl
1,
Jeta Ramadani-Muja
1,
Wolfgang F. Graier
1,2 and
Roland Malli
1,2
1
Gottfried Schatz Research Center, Chair of Molecular Biology and Biochemistry, Medical University of Graz, Neue Stiftingtalstraße 6/6, 8010 Graz, Austria
2
BioTechMed Graz, Mozartgasse 12/II, 8010 Graz, Austria
*
Author to whom correspondence should be addressed.
Submission received: 30 April 2019 / Revised: 17 May 2019 / Accepted: 21 May 2019 / Published: 22 May 2019
(This article belongs to the Special Issue Emerging Trends in Metal Biochemistry)

Abstract

:
Essential biochemical reactions and processes within living organisms are coupled to subcellular fluctuations of metal ions. Disturbances in cellular metal ion homeostasis are frequently associated with pathological alterations, including neurotoxicity causing neurodegeneration, as well as metabolic disorders or cancer. Considering these important aspects of the cellular metal ion homeostasis in health and disease, measurements of subcellular ion signals are of broad scientific interest. The investigation of the cellular ion homeostasis using classical biochemical methods is quite difficult, often even not feasible or requires large cell numbers. Here, we report of genetically encoded fluorescent probes that enable the visualization of metal ion dynamics within individual living cells and their organelles with high temporal and spatial resolution. Generally, these probes consist of specific ion binding domains fused to fluorescent protein(s), altering their fluorescent properties upon ion binding. This review focuses on the functionality and potential of these genetically encoded fluorescent tools which enable monitoring (sub)cellular concentrations of alkali metals such as K+, alkaline earth metals including Mg2+ and Ca2+, and transition metals including Cu+/Cu2+ and Zn2+. Moreover, we discuss possible approaches for the development and application of novel metal ion biosensors for Fe2+/Fe3+, Mn2+ and Na+.

Graphical Abstract

1. Introduction

Vital cells tightly control their (sub)cellular ion distributions [1]. Alterations of the intracellular ion homeostasis are associated with severe dysfunctions and pathologies. Frequently, neurodegenerative processes, cancerous alterations and systemic diseases are linked to changes in the intracellular ion composition [2,3,4,5,6,7]. Especially the distribution of metal ions within cells is of great importance, as they can act as second messengers, modulate signaling pathways or directly affect enzyme functions due to their (in)direct involvement in enzyme-substrate interactions [2,8,9,10,11,12]. However, our current knowledge about intracellular ion concentrations in different organelles is quite limited mainly due to the lack of suitable quantification methods.
Besides the importance of an intact intracellular ion homeostasis, the extracellular ion composition is also of fundamental importance to maintain cellular physiology (Figure 1) [13]. Cells communicate with each other via their (local) environment, where minor changes of the ion composition within the extracellular fluid (ECF) can affect cell activities and even promote the development of severe diseases [4,14,15].
Thereby, especially in the surrounding of excitable cells or cells undergoing cell death, the ion composition of the ECF is drastically affected [4,16,17]. Notably, ion changes in the ECF impact the intracellular ion homeostasis, which might lead to compensatory mechanisms of cells to maintain their intracellular ion composition [18,19]. As an example, traumatic brain injuries are associated with extracellular potassium ion concentration ([K+]) alterations in the brain. These extracellular potassium ion (K+) fluctuations have been linked with a poor survival of these patients, due to an increase of the intracranial pressure, caused by either cell swelling or altered vasoreactivity in cerebral blood vessels [20]. Moreover, an intake of angiotensin-converting-enzyme (ACE) inhibitors by patients suffering from renal insufficiency was linked to the development of severe hyperkalemia, which in turn promotes the development of cardiac arrhythmias due to cardiac-myocyte depolarization [21,22,23].
Although measurements of the extracellular ion concentrations are feasible using clinic-chemical detection methods, their intra- and subcellular distributions and dynamics upon cell stimulation are difficult to capture [24]. The development of reliable detection methods for intracellular metal ions is inevitable to gain our understanding of the mechanisms ultimately causing diseases and may serve as an essential basis for the development and improvement of medical therapies [25].
Valuable tools for the performance of such measurements represent genetically encoded, fluorescent biosensors. The discovery and isolation of the green fluorescent protein (GFP) in 1962 by Shimomura et al. paved the way for the development of GFP-based probes and tools, which have boosted our understanding of cell biology as well as molecular and cellular (patho)physiology [26,27]. Nowadays, numerous differently colored fluorescent protein (FP) variants have been engineered [28]. The first successful design of a so-called genetically encoded probe (GEP) was achieved in 1997 by Miyawaki et al., who exploited Förster resonance energy transfer (FRET) occurring between a yellow and a cyan fluorescent protein (YFP, CFP) variant, fused to specific calcium ion (Ca2+) binding domains [27]. These probes were based on a conformational rearrangement upon Ca2+ binding, yielding an increase in FRET, while the donor fluorescence, i.e., CFP, decreases [27]. The development of this sophisticated probes, referred to as “Cameleons”, opened the door for many further FRET-based indicators that allow real-time recordings of cell signaling events [29,30,31,32]. Nowadays, a huge variety of GEPs is available, especially for metal ions including protein-based fluorescent sensors for Ca2+, copper ions (Cu+/Cu2+), magnesium ions (Mg2+), K+ and zinc ions (Zn2+) [30,33,34,35,36].
In principle, these probes mostly consist of CFP-YFP (Figure 2a) or GFP- red fluorescent protein (RFP) (Figure 2b), fused to specific ion binding domains [27,30,32]. Ion binding to the ion sensitive domain within the probe changes the distance of the two FPs located at the termini of the binding domain, which leads to alterations of the fluorescent properties (Figure 2) [37]. This design allows dynamic and ratiometric read-outs of cellular ion fluctuations, as upon excitation of the donor FP, two emission wavelengths are recorded simultaneously, showing opposite fluorescence alterations depending on the ion concentration (Figure 2c,d). The FRET/donor fluorescence ratio is directly proportional to the concentration of the ion (Figure 2c,d) and allows the investigation of intracellular ion dynamics with high spatial and temporal resolution upon cell treatment (Figure 3) [37].
Besides these FRET-based sensors, a broad range of different probes consisting of only one FP have been developed. These indicators feature some advantages over their FRET-based ancestors, including narrower spectral properties, small size and a huge dynamic range [38,39,40,41]. A detailed review focusing on diverse classes of genetically encoded probes is available [42]. Nonetheless, until today FRET-based probes represent the gold standard of GEPs, due to their ratiometric, and thus, easily quantifiable read-out [43]. However, single FP-based probes represent valuable tools to perform multicolor and –parameter imaging on the single cell level, often in combination with FRET-based probes [30,44]. Nevertheless, such measurements require a sophisticated experimental setup to properly distinguish and separate the spectral features of different probes. Several manuscripts focusing on the usage of genetically encoded probes from a rather technical point of view are available [42,45,46].
Notably, a simple fusion with an approved localization domain results in a, mostly, distinct subcellular distribution of these GEPs [47]. In contrast to most chemical ion indicators, the targeting of biosensors allows the visualization and quantification of ion signals on the subcellular level [30,32,48].
Here we review a selection of frequently used FRET-based GEPs, suitable for the high-resolution investigation of physiologically important, (sub)cellular alkali metals such as K+, alkaline earth metals including Mg2+ and Ca2+, and transition metal ions such as Cu+/Cu2+ and Zn2+. Additionally, we will focus on novel strategies for the design of probes, which might be suitable for the visualization of sodium ions (Na+), iron ions (Fe2+/Fe3+) and manganese ions (Mn2+) on the single cell level.

2. Genetically Encoded Fluorescent Probes for Imaging the Alkali Metal Ion K+, Highly Desired and Freshly Introduced

Living cells maintain a steep K+ gradient across their plasma membrane, while K+ represents the most abundant intracellular cation [13]. Besides its fundamental involvement in the generation of the electrochemical gradient important for many cell types in order to remain excitable, K+ is essentially involved in the regulation of the cell volume, endo- and exocytosis or the prevention of apoptosis [49,50,51,52,53]. In addition, K+ is a central cofactor of several enzymes. The pyruvate kinase, the pyruvate dehydrogenase kinase, the diol- and the glycerol dehydratase or the ribokinase are only some of many examples, pointing to K+ as a master regulator of metabolic pathways [54,55,56,57,58,59]. Despite the important role of K+ in cell physiology, the development of K+ sensitive probes has also been hampered by the lack of specific K+ binding domains, suitable for the use within GEPs. Thus, several chemical fluorescent K+ indicators, based on crown ethers, applicable for the visualization of intracellular [K+] have been developed [60,61,62]. However, these K+ binding probes such as PBFI and Asante Potassium Green-1 (APG-1) suffer from poor specificity for K+ over Na+, require phototoxic excitation with UV-light or absolute quantifications of [K+] are difficult [60,61].
Recently, Ashraf and coworkers unraveled the function of a K+ binding protein, shortly Kbp (K+ binding protein), formerly YgaU, a protein of unknown function until then, found in Escherichia coli (E. coli) [63]. The expression of Kbp, consisting of a bacterial OsmY and nodulation (BON) domain and a lysine motife (LysM), was known to be promoted by the bacterial RNA polymerase sigma S (RpoS) regulon [64]. In principle, genes underlying the RpoS regulon in E. coli are associated with the stress response, thereby acting pro- and retroactive [65]. Within their study, Ashraf et al. demonstrated that Kbp represents a K+ binding protein in these bacteria, important to enable bacterial growth at extracellular [K+] above 1 mM [63]. Upon K+ binding, Kbp undergoes a huge conformational rearrangement, making it ideal for the design of GEPs for K+ sensing and imaging [30,63].
Based on Kbp, we and others have recently developed K+ sensitive FRET-based GEPs referred to as Genetically Encoded Potassium Ion Indicators (GEPIIs) [30,66]. The GEPIIs proved suitable for the determination of extra- and intracellular [K+]. Thereby, the GEPII variant containing the wild type Kbp, referred to as GEPII 1.0, showed a dissociation rate constant (KD) of ~ 0.8 mM when expressed in cells, which is suitable for the quantification of extracellular [K+] [30,67]. However, the high affinity of GEPII 1.0 is far from physiologic intracellular [K+] of ~ 140 mM [68]. Thus, for being able to measure intracellular K+ fluctuations, different approaches to desensitize the probe for K+ were tested, including the introduction of linkers of variable length between the BON and the LysM domain or the insertion of point mutations within the domains, respectively. The introduction of point mutations resulted in probes referred to as low charge (lc-) BON GEPII 1.0 and lc-LysM GEPII 1.0, which either totally lost their K+ sensitivity or showed much lower affinity for K+ than GEPII 1.0 [30].
Expression of lc-LysM GEPII 1.0 within various organelles of several cell lines allowed the quantification of subcellular [K+] within the cytosol, mitochondria and nucleus of these cells. Interestingly, cell line specific differences of the nuclear/cytosolic as well as mitochondrial/cytosolic [K+] have been described. These differences might represent a hallmark of cancerous alterations [30]. However, direct evidences for correlations of cancerous malignancies and the distribution of intracellular [K+] are still missing. Nonetheless, GEPIIs will significantly contribute to our understanding of the K+ distribution, its dynamics and the role of K+ in health and disease. In particular, K+ has recently been demonstrated to be involved in the onset and progression of tumor growth [4]. Direct evidences for an increased intracellular [K+] ([K+]i) are still missing, probably due to the lack of applicable detection methods for K+. GEPIIs could therefore represent valuable tools to correlate the [K+]i with the degree of suppression of T-cell receptor (TCR) activation.
Besides the quantification of [K+]i, such GEPIIs represent applicable tools for the visualization of K+ fluctuations over time. Recently, these FRET-based GEPs were simultaneously expressed in single cells with genetically encoded, single FP-based, red fluorescent Ca2+ indicators [41]. Cells expressing the probes were analyzed in response to cell depolarization (Figure 4a). Such experiments allowed the correlation of [K+]i with intracellular Ca2+ alterations upon cell depolarization (Figure 4b) [30].
These genetically encoded K+ indicators have great potential to deepen our knowledge of the cellular K+ homeostasis on the level of different organelles. Furthermore, understanding [K+] distributions under conditions of health and disease may allow the development of novel therapeutic approaches. A selection of fluorescent K+ indicators suitable for intracellular K+ imaging is given in Table 1.

3. Genetically Encoded Fluorescent Probes for Alkaline Earth Metal Ions

3.1. Genetically Encoded Mg2+ Indicators, Sophisticated Tools, Rarely Applied

Besides K+ as the most abundant intracellular monovalent cation, Mg2+ represents the most abundant intracellular divalent cation [69]. Mg2+ is fundamentally involved in a huge variety of essential cellular processes including transcription, cell replication, energy metabolism or the regulation of specific ion channels [70,71,72,73]. Besides, Mg2+ is a central cofactor of many enzymes and regulates especially the activity of glycolytic enzymes in the chelated form of MgATP2− [74,75,76]. In addition, Mg2+ is considered to be involved in the intracellular signal transduction as a second messenger [77]. The total intracellular Mg2+ concentration ([Mg2+]i) ranges from 17–20 mM, while the free Mg2+ concentration was estimated to be around 1 mM [69]. Various hereditary disorders such as diminished kidney function, renal failure, muscle spasms and seizures have been linked to disorders in the Mg2+ homeostasis [78,79,80,81]. Additionally, the onset of cellular senescence and thus ageing-related diseases such as diabetes or cardiovascular disease have been linked to Mg2+ homeostasis alterations [82,83,84].
Considering the importance of intracellular Mg2+, scientists early aimed for the development of Mg2+ sensitive fluorescent indicators such as Mag-fura-2 [85]. However, the first FRET-based GEP sensitive for Mg2+ was introduced by Lindenburg et al. not earlier than 2013 [34]. They developed a Mg2+ sensitive GEP, based on the human centrin isoform HsCen3 [34]. In principle, centrins represent proteins associated with the centrosome of cells, important for centrosome duplication and separation [86]. It has been demonstrated earlier that HsCen3 has a mixed Ca2+ and Mg2+ binding site possessing high affinity at its N-terminus, while a second Ca2+ specific binding domain located at the C-terminus shows low Ca2+ specific affinity [87].
In the study of Lindenburg et al., truncation of the protein and the usage of the N-terminal fragment within an FP FRET-pair led to the construction of a probe sensitive for Mg2+, referred to as MagFRET. In the apo state, HsCen3 forms a molten globule, which folds into a compact EF-hand like structure upon Mg2+ fixation. The introduction of single point mutations within the EF-hand like motifs led to the construction of a huge variety of Mg2+ sensitive indicators, showing different Mg2+ affinities and different specificities over Ca2+ binding [34]. However, these probes have only rarely been used for the investigation of intracellular Mg2+ fluctuations on the level of individual cells.
Another approach of the development of genetically encoded Mg2+ indicators represents the magnesium ratiometric indicator for optical imaging (MARIO) [88]. The function of the FRET-based probe MARIO is based on the Mg2+ transporter CorA found in E. coli, undergoing a conformational rearrangement upon Mg2+ binding [89]. In bacteria, CorA mediates Mg2+ influxes and effluxes into and from the cells [90]. CorA localizes within the membrane of E. coli [90]. Nonetheless, structural data of CorA seem controversial, as older studies suggest that the majority of the protein is localized in the periplasm, with a small C-terminal part in the cytosol, while recent studies imply only a cytoplasmic localization [89,90]. For the generation of MARIO the original protein of CorA was modified by deleting the N-terminal region of the protein, as structural data suggested that this region is not involved in Mg2+ binding [88,89]. Additionally, polar-charged amino acids were replaced and random mutations were introduced. Performing these modifications of CorA and usage of the protein within a FP FRET-pair yielded a probe showing a lower KD for Ca2+ (6.2 mM) than for Mg2+ (7.2 mM). Nonetheless, the Ca2+ sensitivity of this sensor is far from the physiological cytoplasmic range [91]. Using MARIO, the authors reported an increase in the free cytosolic Mg2+ concentration ([Mg2+]), which contributes to mitotic chromosome condensation [88]. However, further applications of MagFRET or MARIO have not been performed until today, thus, insights into (patho-)physiological alterations of the intra- and especially subcellular Mg2+ homeostasis remain mainly elusive. A selection of FRET-based genetically encoded Mg2+ indicators suitable for intracellular Mg2+ imaging is given in Table 2.

3.2. Genetically Encoded Ca2+ Indicators, A Huge Variety for An Ion with Versatile Roles

Ca2+ represents a central second messenger in living cells and is, thus, involved in countless signaling pathways [92]. Besides its involvement in the regulation of enzymatic activities as well as ion channels and ion pumps, Ca2+ fundamentally contributes to the regulation of gene transcription [93,94,95,96,97]. Additionally, Ca2+ participates in the contraction of all muscle cells, neurotransmitter and hormone release or fertilization [98,99,100,101]. Especially the mitochondrial and ER Ca2+ homeostasis and the Ca2+ concentration ([Ca2+]) within the organelles are tightly regulated [102,103,104]. The ER represents an intracellular Ca2+ storage with [Ca2+] up to 1 mM, which is important for the ER in order to retain its structure and allow protein folding by Ca2+ dependent chaperones such as BiP or the protein disulfide-isomerase (PDI) [105,106,107]. Ca2+ transfer from the ER towards the mitochondrial matrix is necessary to ensure physiological cell bioenergetics [104,108]. The mitochondrial matrix features a strongly negative membrane potential, thus, Ca2+ would rapidly accumulate within the matrix without the strict Ca2+ handling of the organelle, further leading to the induction of apoptosis, and finally cell death [109,110,111]. Ca2+ fluxes into the mitochondrial matrix are, thus, highly regulated and achieved via a whole protein ensemble, especially including the mitochondrial calcium uniporter (MCU), the mitochondrial calcium uptake 1 and 2 (MICU1/2) and the essential MCU regulator (EMRE). Detailed reviews about the regulation of mitochondrial Ca2+ uptake and homeostasis are available [102,112,113].
Considering the importance of Ca2+ as a ubiquitous intracellular messenger, the variety of Ca2+ sensitive sensors represents likely the most extensive one of all available indicators. Until today, one of the most prominent designs of a Ca2+ indicator suitable for the visualization of global intracellular Ca2+ alterations represents Fura-2, a ratiometric chemical fluorescent Ca2+ sensitive dye [114]. However, as all chemical fluorescent indicators, Fura-2 features some severe disadvantages in comparison to GEPs, including phototoxicity due to excitations with UV-light, difficult subcellular targeting and poor applicability in living animals. Thus, scientists aimed for the development of genetically encoded FRET-based GEPs. The first successful design of a FRET-based genetically encoded Ca2+ indicator succeeded in 1997 by Miyawaki et al., who introduced the first genetically encoded sensor of all time and paved the way for countless further developments [27]. By the fusion of a CFP and a YFP variant to calmodulin (CaM), a Ca2+ binding protein, and the myosin light chain kinase M13, a protein interacting with Ca2+ bound CaM, they yielded functional Ca2+ sensors referred to as Cameleons [27]. Nowadays, a huge variety of these Cameleons is available. Cameleons with different Ca2+ sensitivities for different cellular organelles with organelle specific [Ca2+] have been successfully generated [27,32,48,115,116,117,118,119,120,121]. The most frequently used probes thereby contain Ca2+ sensitive domains referred to as D1, D2, D3 or D4, which represent mutated variants of CaM and M13, respectively [27,32,48,115,116,117,118,119,120,121]. Using these domains, the visualization of intracellular relevant [Ca2+] became possible within the mitochondria, the nucleus, the cytosol, the ER, the Golgi apparatus or endosomes of single living cells, all possessing different [Ca2+] [27,32,33,48,115,116,117,118,119,120,121]. Besides FRET-based probes consisting of CFP-YFP as a FRET-pair, probes containing GFP and RFP variants have been developed. Such red-shifted probes allow the performance of simultaneous measurements on the single cell level with, for example, Fura-2 loaded cells and the probe expressed in the ER (Figure 5a) or mitochondria (Figure 5b). Such analysis of Ca2+ within the cytosol and organelles of single living cells over time allows the correlation of subcellular Ca2+ dynamics (Figure 5c,d) [32,48].
Considering the construction of these GEPs sensitive for Ca2+ based on the human variants of CaM and M13, it is difficult to claim whether these probes may interfere with, or promote, intracellular signaling cascades. Therefore, Ca2+ sensitive, FRET-based GEPs using Ca2+ binding domains of non-mammalian origin were developed. Famous examples of such Ca2+ sensitive probes are based on chicken skeletal muscle troponin C (TnC) located within the FP FRET-pair, bearing the advantage of using Ca2+ sensitive domains of non-mammalian origin [122,123,124]. TnC is a Ca2+ binding protein found in the thin filament of skeleton muscle cells. Upon Ca2+ binding to TnC, muscle contraction is initiated due to the change of the proteins conformation which is Ca2+ dependent [125]. Additionally, Ca2+ sensitive probes not directly binding Ca2+, but measuring the Ca2+ bound form of calreticulin, the major Ca2+ binding protein in the lumen of the ER, have been developed [126]. These indicators feature the advantage of being not directly Ca2+ buffering, but measuring an endogenous, Ca2+ bound protein. Using these probes, however, showed similar Ca2+ kinetics within the ER upon cell stimulation in comparison to other frequently used ER targeted, Ca2+ sensitive probes [32,126].
Besides the availability of a large variety of FRET-based indicators for Ca2+, Ca2+ indicators consisting of only one FP variant have been developed [41,44,127]. These single FP-based indicators can function either in an intensiometric manner where only one fluorescence signal is recorded, or allow a ratiometric read-out, indicating that, dependent on the [Ca2+], either two excitation maxima at constant emission, or two emission maxima at constant excitation allow the absolute quantification of [Ca2+]. These probes represent valuable tools for multi-channel and multi-parameter imaging due to their narrow fluorescence properties [30,41,44]. A detailed review focusing, amongst others, on single FP-based genetically encoded Ca2+ indicators is available [128]. A selection of FRET-based genetically encoded Ca2+ indicators suitable for intracellular Ca2+ imaging is given in Table 3.

4. Genetically Encoded Fluorescent Probes for Transition Metal Ions

4.1. Genetically Encoded Cu+/Cu2+ Indicators, Highly Sensitive Tools for Low Concentrated Ions

Copper ions (Cu+/Cu2+) represent an essential element for life due to their fundamental role as cofactors in protein functions [12]. The reversibility of the oxidation and reduction of Cu+ and Cu2+ make the function of this metal ion versatile. The cellular Cu+/Cu2+ homeostasis has to be tightly regulated, as an excess of this ion is associated with severe cellular dysfunctions and pathologies [130]. Especially diseases such as Wilson’s and Menkes disease are linked to dysregulations of the Cu+/Cu2+ uptake and are prominent examples of Cu+/Cu2+ associated pathologies [131,132]. Free Cu+/Cu2+ catalyze Fenton-type reactions, thereby producing ROS [133]. Thus it is not surprising that elevated levels of Cu+/Cu2+ are highly toxic and dysregulations of the Cu+/Cu2+ homeostasis lead to cardiovascular disorders and neurodegeneration [134,135,136].
In principle, Cu+/Cu2+ has high affinity to most ligands, thus the kinetically labile intracellular Cu+/Cu2+ needs to be buffered to low free concentrations to prevent its binding to other metal ion binding sites [137,138]. Therefore, many intracellular ligands for Cu+/Cu2+ exist, preventing unspecific binding [138]. As a result, only high-affinity Cu+/Cu2+ binding sites are able to compete for cellular Cu+/Cu2+, generating the Cu+/Cu2+ specificity of such proteins [36,139]. Due to the intracellular Cu+/Cu2+ buffering, it is more meaningful to consider the global Cu+/Cu2+ availability and the effective intracellular concentration, rather than the free ion concentration [36]. Several small chemical fluorescent indicators applicable for the detection of Cu+ have been developed recently, while only a few Cu+ sensitive GEPs are available [140].
One approach for the detection of Cu+ within single living cells was published in 2010 by Wegner et al., establishing a FRET-based GEP suitable for intracellular Cu+ measurements [36]. The probe referred to as Amt1-FRET enables Cu+ imaging on the single cell level. As the name suggests, the sensor is based on Amt1, a protein from Candida glabrata (C. glabrata), flanked by a CFP and a YFP variant. Only amino acids 36-110 from the whole length protein were used for the design of the probe, as these amino acids represent the actual Cu+ binding structure. Upon Cu+ binding to Amt1, the probe undergoes a conformational rearrangement, yielding increased FRET and decreased donor fluorescence [36]. In C. glabrata, Amt1 represents a Cu+ binding transcription factor, inducing the expression of genes involved in the detoxification of elevated intracellular Cu+ levels [141]. Interestingly, the probe was not only sensitive for Cu+, but had a similar affinity for silver ions (Ag+) and a much lower affinity for Zn2+. Using Amt1-FRET the authors demonstrated an increase in free intracellular Cu+ upon incubation of cells in copper sulfate containing media, which was reversible upon addition of a Cu+ ligand [36]. In addition, Wegner et al. published two FRET-based GEPs in 2011 and demonstrated their suitability for Cu+ imaging in yeast [142]. The probes referred to as Ace1-FRET and Mac1-FRET are designed in analogy to Amt1-FRET and are based on Cu+ dependent transcription factors, also derived from yeast. Ace1 represents a transcription factor activating the expression of Cu+ detoxification and storage genes, while Mac1 senses the lower Cu+ level, regulating the expression of genes responsible for Cu+ uptake [142,143,144,145].
Besides Amt1-FRET, AceI-FRET and MacI-FRET, which are suitable for Cu+ measurements in the atto- to zeptomolar range, Yan et al. established a novel FRET-based Cu+ sensor in 2012 [146]. The GEP referred to as PMtb-FRET is based on the conformational change of CDC 1551 derived from Mycobacterium tuberculosis (Mtb) [147]. Mtb is the bacterium responsible for causing tuberculosis, a well-known disease of the lungs, but can also attack other parts of the body, including the brain, spine, kidney and arthrosis [148,149,150,151]. It is believed that especially Cu+-binding domains of Mtb cause its ability to adapt to a huge variety of challenging environments [152]. For their design of a FRET-based probe Yan et al. sub-cloned the amino acid residues 1-162, which represent the actual Cu+-binding domain of CDC 1551, into a CFP-YFP FRET-pair. Upon Cu+-binding to the peptide, FRET-efficiency is increasing, enabling a ratiometric read-out of the sensor. The probe showed higher sensitivity for Cu+ in the zeptomolar range. Similar to Amt1-FRET the sensor responds to other ions including Zn2+, Fe2+ and cadmium ions (Cd2+) [146]. Nonetheless, PMtb-FRET was not tested upon expression within single living cells in their study, hence it remains elusive whether PMtb-FRET represents a tool suitable for the visualization of intracellular Cu+ dynamics.
In general, FRET-based Cu+-sensitive probes have only rarely been applied for intracellular Cu+ measurements, yet their properties and features seem highly promising for the visualization of intra- and subcellular Cu+ dynamics under diverse experimental setups [36,142,146,153]. A selection of FRET-based genetically encoded Cu+ indicators for intracellular Cu+ imaging is given in Table 4.

4.2. Genetically Encoded Zn2+ Indicators, A Broad Palette of Applicable Probes

Intracellular Zn2+ is an important micronutrient for many fundamental cellular processes, including structural functions in DNA-binding proteins, non-redox catalysis as well as neurotransmission [154,155,156]. Interestingly, around 10% of all genes code for Zn2+-binding motifs, demonstrating its importance in cell physiology [157]. Zn2+ homeostasis needs to be tightly regulated on the (sub)cellular level, as low micromolar concentrations of free Zn2+ can be cytotoxic [158]. Still, especially the intracellular Zn2+ homeostasis is poorly understood, making it an interesting target for the design of fluorescent, Zn2+ sensitive GEPs [159]. Notably, FRET-based probes designed for the measurement of intracellular Zn2+ concentrations ([Zn2+)i] need to be highly sensitive and measure Zn2+ concentrations ([Zn2+]) in the nanomolar range.
The first successful design of a genetically encoded, FRET-based indicator suitable for the detection of Zn2+ was achieved by van Dongen et al. in 2006 [35]. The probe is constructed from two different proteins which should interact upon Cu+ binding and consists of cyan and yellow FP variants fused to Atox1 or the fourth domain of copper transporting P-type ATPase ATP7B (WD4), respectively. In principle, Atox1 represents a human copper chaperone protein which targets Cu+ for excretion and incorporation into extracellular Cu+ proteins, while ATP7B is located in the Golgi membrane and represents the actual Cu+ transporter [160,161]. Initially, the probe referred to as CA + WY (CFP-Atox1 + WD4-YFP) was designed to sense Cu+, however, increases in FRET-efficiency appeared very unstable, as DTT could disrupt the Cu+-Atox1-WD4 complex. Interestingly, the probe responded to very low concentrations of Zn2+ of 10−10 M, also in the presence of DTT. Thus, the authors aimed to optimize the Zn2+ sensing behavior of the indicator and mutated the Cu+-binding motifs for known Zn2+-binding consensus sequences [35]. Nonetheless, the probe was not applied for the detection of intracellular Zn2+, probably due to the low dynamic range of the sensor and its lack of specificity for other ions such as cobalt ions (Co2+), plumb ions (Pb2+) and Cd2+.
In 2007, van Dongen et al. optimized their first design of the Zn2+ sensitive probe CA + WY and designed fusion constructs of CA + WY by introducing linkers of different lengths between the two proteins. Insertion of the linkers drastically increased the dynamic range and the sensitivity of the probes referred to as CA-L2-WY containing two, CA-L5-WY containing five and CA-L9-WY containing nine GGSGGS repeats [162].
Dittmer et al. introduced the first FRET-based indicators for the intracellular visualization of Zn2+ in 2009 [163]. The authors generated a FRET-based GEP based on a CFP and a YFP variant, fused to the terminal ends of a canonical Cys2His2 zinc finger, derived from the mammalian transcription factor Zif268, a Zn2+ specific binding protein [163]. Structural data of Zif268 suggested an unstructured conformation in the absence of the metal ion which folds into a compact structure upon Zn2+ binding [164]. In mammals, zinc fingers of the Cys2His2 class recognize a wide variety of different DNA sequences and mostly represent trans regulators of gene expression [165]. Thus, they are fundamentally involved in development, differentiation and oncosuppression. Generally, zinc fingers represent short protein motifs consisting of two or three β-layers and one α-helix. These motifs bind Zn2+ which stabilizes their structure, enabling their function as transcription factors [165]. Besides their interaction with DNA, zinc fingers can also provide protein-protein and RNA-protein interactions [166,167]. However, the usage of the Cys2His2 domain derived from Zif268 within a FP FRET-pair proved suitable for the design of functional Zn2+ sensitive indicators. Moreover, the authors introduced mutations in the Cys2His2 domain and yielded FRET-based constructs containing His4, showing lower Zn2+ affinity but a higher dynamic range, or created a construct containing Ala4, lacking Zn2+ sensitivity. In their study, Dittmar et al. applied their novel probes for the quantification of [Zn2+] within the cytosol and mitochondria of single living cells with and determined [Zn2+] of 180 nM and 680 nM for both localizations, respectively [163].
Nowadays, a huge variety of applicable detection methods for intracellular Zn2+ is available [140]. One interesting approach for the development of Zn2+ sensitive probes represent the sensors referred to as ZinCh [168]. These sensors consist of a CFP and a YFP variant, fused via a flexible linker. The FPs within the probe possess point mutations, directly generating a Zn2+ sensitive pocket, yielding high FRET-efficiency in the presence of Zn2+, while in the absence of Zn2+ FRET-fluorescence is low and CFP fluorescence is increased [168]. Recently, these probes were optimized yielding eZinCh-2, suitable for the determination of intracellular [Zn2+] [169]. Besides the expression in the cytosol of cells, the indicator was targeted to the ER, mitochondria and insulin secreting vesicles, enabling the dynamic visualization of Zn2+ alterations with high spatial and temporal resolution [169]. A selection of FRET-based genetically encoded Zn2+ indicators suitable for intracellular Zn2+ imaging is given in Table 5.

5. Approaches for the Design of Novel Genetically Encoded Na+, Fe2+/Fe3+ and Mn2+ Indicators

Na+ represents the most abundant extracellular cation, involved in the maintenance of the cell membrane potential [172,173]. Na+ is especially important for the induction of action potentials, as it represents the first ion flowing across the membrane upon cell membrane depolarization [174]. In addition, extracellular Na+ concentrations ([Na+]ex) are a master regulator of the osmotic equilibrium in organisms, regulate the blood volume as well as the blood pressure [175]. Furthermore, Na+ functions as an important cofactor of clotting proteases such as thrombin [176].
Our understanding of the actual subcellular Na+ homeostasis and its role in regulating cell physiology and cellular functions remains, however, quite poor. Although small chemical fluorescent indicators sensitive for Na+ are available, these probes do not represent applicable detection methods. These fluorescent chemical Na+ indicators bind Na+ via a crown ether structure within a certain cavity [177,178,179,180,181]. However, most available probes suffer from poor specificity for Na+ over K+, and their affinity for Na+ appears to be dependent on [K+] or is far from the physiologic intracellular Na+ concentration ([Na+]) [177,181,182]. On the other hand, several probes only allow an intensiometric read-out, making quantifications of absolute [Na+] challenging [178,179,180,181,182]. Additionally, all of these chemical probes are difficult to target to subcellular structures such as the endoplasmic reticulum (ER), the nucleus or the Golgi apparatus, although mitochondria localized Na+ indicators are available [182]. Altogether, these clear disadvantages and shortcomings of chemical Na+ indicators demonstrate the urgent need for an applicable detection method of subcellular Na+. One approach to yield such optimized probes might be the development of genetically encoded, FRET-based Na+ indicators. Although several proteins specifically binding Na+ have been described, a development of GEPs for Na+ has mainly been hampered by the lack of Na+ specific binding domains, undergoing a conformational rearrangement upon Na+ binding [183,184,185]. The identification of such a peptide may allow the design of the first GEP sensitive for Na+. In addition, such GEPs will appear highly advantageous over available chemical indicators, as they can be targeted to subcellular structures and organelles, are non-toxic upon cellular expression and can also be applied in living animals [28,47]. Still the probe needs to be specific for Na+ over all other ions, and the Na+ binding should not appear in a [K+] dependent manner.
Iron is fundamentally involved in important cellular functions and is required by all known living organisms [186]. In biology, iron almost exclusively exists in the form of Fe2+ (ferrous state) or Fe3+ (ferric state), although Fe3+ is only poorly soluble at neutral pH in aqueous media [140]. Hence, the cytosolic reduction potential favors Fe2+ over Fe3+, which is involved in ensuring the proper function of Fe2+-requiring enzymes [187]. These enzymes usually contain heme prosthetic groups, important for the catalysis of oxidation reactions [188]. Fe2+ participates in the transport of soluble gases, as it is a central component of hemoglobin and represents an essential constituent of cytochromes and other components of respiratory enzyme systems [189,190,191]. Additionally, inorganic iron is involved in redox reactions, and hence is a fundamental cofactor for enzymes such as the nitrogenase, involved in the synthesis of ammonia from nitrogen and hydrogen, hydrogenases, catalyzing the reversible oxidation of hydrogen, ribonucleotide reductase, reducing ribose to deoxyribose for DNA biosynthesis, or purple acid phosphatase, hydrolyzing phosphate esters [192,193,194,195]. Typically, cells store intracellular iron via ferritin, a Fe2+ binding protein [196].
Besides the importance of Fe2+ for essential biological functions, free Fe2+ participates in Fenton chemistry, promoting the production of reactive oxygen species (ROS) causing cell toxicity [197]. For this reason, the intracellular Fe2+ homeostasis needs to be tightly regulated, in order to balance the absolute cellular requirement on Fe2+ and toxicity induced by excess Fe2+ [198]. According to the importance of Fe2+ for fundamental cellular functions, numerous fluorescent chemical probes suitable for the detection of intracellular Fe2+ and Fe3+ have been developed. In principle, Fe2+/Fe3+ binding to these fluorescent chemical sensors changes their conformation, causing alterations of their fluorescence properties. These indicators either work in a turn-on or turn-off manner, indicating that their fluorescence is increasing or quenched upon Fe2+/Fe3+ binding, however, also ratiometric sensors have been developed [140,199,200,201,202,203].
Mn2+ represents an essential micronutrient for all kingdoms of life [204]. Besides its importance as a cofactor of many critical enzymes, the mitochondrial redox balance, the cellular adaption to oxidative stress, intracellular vesicle shuttling and neuron function are dependent on Mn2+ [205,206,207,208,209]. As with most metal ions, excess of Mn2+ can be toxic for cells [210]. Especially neurotoxicity is linked to Mn2+ accumulations and symptoms similar to those from Parkinson’s disease have been linked to disturbances of the neuronal Mn2+ homeostasis [209]. Considering these manifold roles of Mn2+ in health and disease, scientists developed chemical fluorescent indicators, suitable for Mn2+ imaging on the level of single cells [211]. Upon Mn2+ binding to these probes, their fluorescence is drastically increasing, allowing an intensiometric read-out.
However, the development of Fe2+, Fe3+ and Mn2+ sensitive GEPs is still hampered by the identification of suitable ion binding domains. Interestingly, several Mn2+ binding proteins, especially from bacterial origin, have been described and analyzed [212,213]. Also enzymes requiring and binding Mn2+ in mammals have been identified [214,215]. Similarly, several iron binding sites are known, presenting opportunities for the development of Fe2+, Fe3+ and Mn2+ sensitive probes. However, to our knowledge, these proteins have probably either not been tested within GEPs yet or did not prove suitable for the usage in a respective protein-based fluorescent metal ion indicator.
Ideally, ion sensitive domains used within these GEPs would be highly selective for one ion over all other ions, bind the ion reversibly, repetitively and with fast kinetics. Additionally, the design of FRET-based indicators requires a protein undergoing a change in conformation upon ion binding [37]. However, other approaches such as the usage of ion dependent protein-protein interactions have also been applied and yielded functional probes [35]. One possibility to develop such GEPs applicable for the visualization of Fe2+, Fe3+ or Mn2+ may aim to take advantage of an unspecific ion binding by already existing probes. One could use existing probes which are not specifically binding the desired ion with low affinity and try to modify the binding site by the insertion of mutations, in order to change the ion selectivity. That such an approach may be suitable for the generation of genetically encoded Fe2+/Fe3+ and Mn2+ probes has been recently proven by Koay et al. who modified the specificity of a GEP by the introduction of mutations in an approved Cu+ sensor, thereby generating probes sensitive for Zn2+ [153].

6. Concluding Remarks and Outlook

This review provides an overview of genetically encoded, FRET-based biosensors, allowing the investigation of intracellular ion dynamics and concentrations in living cells with high spatiotemporal resolution. We present selected, commonly used sensors for the most important ions in cell physiology including K+, Mg2+, Ca2+, Cu+/Cu2+ and Zn2+, and address possible approaches which might yield novel GEPs suitable for the visualization of Na+, Mn2+ and Fe2+/Fe3+. The advantages of using GEPs instead of chemical fluorescent indicators are highlighted. Hence, we demonstrate remarkable examples of simultaneous applications of either two different GEPs or GEPs in combination with chemical fluorescent indicators. Both approaches drastically increase the spatiotemporal information of (sub)cellular ion signals. Summing up, we want to offer scientists an overview of GEPs suitable for the visualization of intracellular metal ions and their applications.
However, the visualization of intracellular ion signals using FRET-based probes requires a sophisticated experimental setup, as upon excitation of the donor FP two emission wavelengths have to be recorded simultaneously [42,46,48]. Despite the refined technical requirements to measure ion signals using FRET-based sensors, their application in vivo has been demonstrated. FRET-based K+ sensors proved applicable for extracellular K+ measurements, or FRET-based Ca2+ indicators expressed in neurons were applied for the visualization of intracellular Ca2+ signals [30,216].
The broad spectral properties of the indicators often hamper their combination with other probes for multi-parameter imaging. Multiple applications of FRET-based sensors in combination with single FP-based indicators have been demonstrated [30,32,48]. As FRET-based sensors consist of two FPs, they represent relatively bulky constructs. Frequently, their targeting to subcellular compartments appears difficult e.g., into the matrix of mitochondria. However, it has been demonstrated that the fusion of multiple targeting peptide repeats can facilitate and significantly improve the targeting of these probes, resulting in their distinct subcellular localization [119].
FRET-based indicators represent the gold standard for intracellular ion concentration measurements. Their feature of a ratiometric read-out enables ion quantifications almost independent of the concentration of the probes. Nonetheless, different maturation times as well as partial sensor degradation of donor- and acceptor FPs used within FRET-based probes may have an effect on the quantification of ions using these indicators [39]. Additionally, these probes cannot be applied under hypoxic conditions, as molecular oxygen (O2) is required for the maturation of the FP chromophore [217].

Funding

The research of the authors is funded by the Ph.D. program Molecular Medicine (MOLMED) of the Medical University of Graz, by Nikon Austria within the Nikon-Center of Excellence, Graz, the Austrian Science Fund (FWF) with projects P28529-B27 and I3716-B27 to R.M., the Doctoral Program Metabolic and Cardiovascular Disease (DK-W1226). The Nikon Center of Excellence, Graz, is supported by the Austrian infrastructure program 2013/2014, Nikon Austria Inc., and BioTechMed, Graz.

Acknowledgments

The authors acknowledge the scientific advisory board of Next Generation Fluorescence Imaging (NGFI) GmbH (http://www.ngfi.eu), a spin-off company of the Medical University of Graz and the technical support of Anna Schreilechner and Sandra Blass.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

[Ca2+]Calcium ion concentration
[K+]Potassium ion concentration
[K+]iIntracellular potassium ion concentration
[Mg2+]Magnesium ion concentration
[Mg2+]iIntracellular magnesium ion concentration
[Na+]Sodium ion concentration
[Na+]exExtracellular sodium ion concentration
[Zn2+]Zinc ion concentration
[Zn2+]iIntracellular zinc ion concentration
ACEAngiotensin-converting enzyme
Ag+Silver ion
APG-1Asante potassium green-1
BONBacterial OsmY and nodulation
Ca2+Calcium ion
CaMCalmodulin
Cd2+Cadmium ion
CFPCyan fluorescent protein
Co2+Cobalt ion
Cu+/Cu2+Copper ions
ECFExtracellular fluid
EMREEssential mitochondrial calcium uniporter regulator
EREndoplasmic reticulum
Fe2+/Fe3+Iron ions
FPFluorescent protein
FRETFörster resonance energy transfer
GEPGenetically encoded probe
GEPIIGenetically encoded potassium ion indicator
GFPGreen fluorescent protein
K+Potassium ion
KbpPotassium ion binding protein
KDDissociation rate constant
lc-Low charge
LysMLysine motife
MARIOMagnesium ratiometric indicator for optical imaging
MCUMitochondrial calcium uniporter
Mg2+Magnesium ion
MICU1/2Mitochondrial calcium uptake 1 and 2
Mn2+Manganese ion
Na+Sodium ion
Pb2+Plumb ion
RFPRed fluorescent protein
ROSReactive oxygen species
RpoSRNA polymerase sigma S
TCRT-cell receptor
TnCTroponin C
YFPYellow fluorescent protein
Zn2+Zinc ion

References

  1. Dubyak, G.R. Ion homeostasis, channels, and transporters: An update on cellular mechanisms. Adv. Physiol. Educ. 2004, 28, 143–154. [Google Scholar] [CrossRef] [PubMed]
  2. Stelmashook, E.V.; Isaev, N.K.; Genrikhs, E.E.; Amelkina, G.A.; Khaspekov, L.G.; Skrebitsky, V.G.; Illarioshkin, S.N. Role of zinc and copper ions in the pathogenetic mechanisms of Alzheimer’s and Parkinson’s diseases. Biochem. Mosc. 2014, 79, 391–396. [Google Scholar] [CrossRef]
  3. Jansson, B. Potassium, sodium, and cancer: A review. J. Env. Pathol. Toxicol. Oncol. 1996, 15, 65–73. [Google Scholar]
  4. Eil, R.; Vodnala, S.K.; Clever, D.; Klebanoff, C.A.; Sukumar, M.; Pan, J.H.; Palmer, D.C.; Gros, A.; Yamamoto, T.N.; Patel, S.J.; et al. Ionic immune suppression within the tumour microenvironment limits T cell effector function. Nature 2016, 537, 539–543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kolobkova, Y.A.; Vigont, V.A.; Shalygin, A.V.; Kaznacheyeva, E.V. Huntington’s Disease: Calcium Dyshomeostasis and Pathology Models. ACTA Nat. 2017, 9, 34–46. [Google Scholar] [CrossRef]
  6. Surmeier, D.J.; Schumacker, P.T.; Guzman, J.D.; Ilijic, E.; Yang, B.; Zampese, E. Calcium and Parkinson’s disease. Biochem. Biophys. Res. Commun. 2017, 483, 1013–1019. [Google Scholar] [CrossRef]
  7. Fuentealba, I.C.; Aburto, E.M. Animal models of copper-associated liver disease. Comp. Hepatol. 2003, 2, 5. [Google Scholar] [CrossRef]
  8. Blaustein, M.P. Intracellular calcium as a second messenger. In Calcium in Biological Systems; Rubin, R.P., Weiss, G.B., Putney, J.W., Eds.; Springer: Boston, MA, USA, 1985; pp. 23–33. ISBN 978-1-4612-9453-5. [Google Scholar]
  9. Page, M.J.; Di Cera, E. Role of Na+ and K+ in Enzyme Function. Physiol. Rev. 2006, 86, 1049–1092. [Google Scholar] [CrossRef] [PubMed]
  10. Pilchova, I.; Klacanova, K.; Tatarkova, Z.; Kaplan, P.; Racay, P. The Involvement of Mg2+ in Regulation of Cellular and Mitochondrial Functions. Oxidative Med. Cell. Longev. 2017, 2017, 1–8. [Google Scholar] [CrossRef]
  11. Gohara, D.W.; Di Cera, E. Molecular Mechanisms of Enzyme Activation by Monovalent Cations. J. Biol. Chem. 2016, 291, 20840–20848. [Google Scholar] [CrossRef] [Green Version]
  12. Festa, R.A.; Thiele, D.J. Copper: An essential metal in biology. Curr. Biol. 2011, 21, R877–R883. [Google Scholar] [CrossRef] [Green Version]
  13. Terry, J. The major electrolytes: Sodium, potassium, and chloride. J. Intraven. Nurs. 1994, 17, 240–247. [Google Scholar]
  14. Díaz-Soto, G.; Rocher, A.; García-Rodríguez, C.; Núñez, L.; Villalobos, C. The Calcium-Sensing Receptor in Health and Disease. Int. Rev. Cell Mol. Biol. 2016, 327, 321–369. [Google Scholar]
  15. Bouschet, T.; Henley, J.M. Calcium as an extracellular signalling molecule: Perspectives on the Calcium Sensing Receptor in the brain. Comptes Rendus Biol. 2005, 328, 691–700. [Google Scholar] [CrossRef]
  16. Harjes, D.I.; Dubach, J.M.; Rosenzweig, A.; Das, S.; Clark, H.A. Ion-Selective Optodes Measure Extracellular Potassium Flux in Excitable Cells. Macromol. Rapid Commun. 2009, 31, 217–221. [Google Scholar] [CrossRef] [Green Version]
  17. Gurusamy, D.; Clever, D.; Eil, R.; Restifo, N.P. Novel “Elements” of Immune Suppression within the Tumor Microenvironment. Cancer Immunol. Res. 2017, 5, 426–433. [Google Scholar] [CrossRef]
  18. Hrabcová, D.; Pásek, M.; Šimurda, J.; Christé, G. Effect of Ion Concentration Changes in the Limited Extracellular Spaces on Sarcolemmal Ion Transport and Ca2+ Turnover in a Model of Human Ventricular Cardiomyocyte. Int. J. Mol. Sci. 2013, 14, 24271–24292. [Google Scholar] [CrossRef]
  19. Walz, W.; Hertz, L. Intracellular ion changes of astrocytes in response to extracellular potassium. J. Neurosci. Res. 1983, 10, 411–423. [Google Scholar] [CrossRef]
  20. Reinert, M.; Khaldi, A.; Zauner, A.; Doppenberg, E.; Choi, S.; Bullock, R. High level of extracellular potassium and its correlates after severe human head injury: Relationship to high intracranial pressure. J. Neurosurg. 2000, 93, 800–807. [Google Scholar] [CrossRef]
  21. Weiss, J.N.; Qu, Z.; Shivkumar, K. Electrophysiology of Hypokalemia and Hyperkalemia. Circ. Arrhythm Electrophysiol. 2017, 10, e004667. [Google Scholar] [CrossRef]
  22. Espinel, E.; Joven, J.; Gil, I.; Suñé, P.; Renedo, B.; Fort, J.; Serón, D. Risk of hyperkalemia in patients with moderate chronic kidney disease initiating angiotensin converting enzyme inhibitors or angiotensin receptor blockers: A randomized study. BMC Res. Notes 2013, 6, 306. [Google Scholar] [CrossRef]
  23. Raebel, M.A. Hyperkalemia Associated with Use of Angiotensin-Converting Enzyme Inhibitors and Angiotensin Receptor Blockers: Hyperkalemia with ACEI and ARB. Cardiovasc. Therap. 2012, 30, e156–e166. [Google Scholar] [CrossRef]
  24. Luxardi, G.; Reid, B.; Ferreira, F.; Maillard, P.; Zhao, M. Measurement of Extracellular Ion Fluxes Using the Ion-selective Self-referencing Microelectrode Technique. J. Vis. Exp. 2015, 99, 52782. [Google Scholar] [CrossRef]
  25. Tian, L.; Looger, L.L. Genetically encoded fluorescent sensors for studying healthy and diseased nervous systems. Drug Discov. Today Dis. Models. 2008, 5, 27–35. [Google Scholar] [CrossRef] [Green Version]
  26. Shimomura, O.; Johnson, F.H.; Saiga, Y. Extraction, Purification and Properties of Aequorin, a Bioluminescent Protein from the Luminous Hydromedusan, Aequorea. J. Cell. Comp. Physiol. 1962, 59, 223–239. [Google Scholar] [CrossRef]
  27. Miyawaki, A.; Llopis, J.; Heim, R.; McCaffery, J.M.; Adams, J.A.; Ikura, M.; Tsien, R.Y. Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin. Nature 1997, 388, 882–887. [Google Scholar] [CrossRef]
  28. Shaner, N.C.; Steinbach, P.A.; Tsien, R.Y. A guide to choosing fluorescent proteins. Nat. Methods 2005, 2, 905–909. [Google Scholar] [CrossRef]
  29. Imamura, H.; Huynh Nhat, K.P.; Togawa, H.; Saito, K.; Iino, R.; Kato-Yamada, Y.; Nagai, T.; Noji, H. Visualization of ATP levels inside single living cells with fluorescence resonance energy transfer-based genetically encoded indicators. Proc. Natl. Acad. Sci. USA 2009, 106, 15651–15656. [Google Scholar] [CrossRef] [Green Version]
  30. Bischof, H.; Rehberg, M.; Stryeck, S.; Artinger, K.; Eroglu, E.; Waldeck-Weiermair, M.; Gottschalk, B.; Rost, R.; Deak, A.T.; Niedrist, T.; et al. Novel genetically encoded fluorescent probes enable real-time detection of potassium in vitro and in vivo. Nat. Commun 2017, 8, 1422. [Google Scholar] [CrossRef]
  31. Burgstaller, S.; Bischof, H.; Gensch, T.; Stryeck, S.; Gottschalk, B.; Ramadani-Muja, J.; Eroglu, E.; Rost, R.; Balfanz, S.; Baumann, A.; et al. pH-Lemon, a Fluorescent Protein-Based pH Reporter for Acidic Compartments. ACS Sens. 2019, 4, 883–891. [Google Scholar] [CrossRef] [Green Version]
  32. Waldeck-Weiermair, M.; Bischof, H.; Blass, S.; Deak, A.; Klec, C.; Graier, T.; Roller, C.; Rost, R.; Eroglu, E.; Gottschalk, B.; et al. Generation of Red-Shifted Cameleons for Imaging Ca2+ Dynamics of the Endoplasmic Reticulum. Sensors 2015, 15, 13052–13068. [Google Scholar] [CrossRef]
  33. Whitaker, M. Genetically encoded probes for measurement of intracellular calcium. Methods Cell Biol. 2010, 99, 153–182. [Google Scholar]
  34. Lindenburg, L.H.; Vinkenborg, J.L.; Oortwijn, J.; Aper, S.J.A.; Merkx, M. MagFRET: The First Genetically Encoded Fluorescent Mg2+ Sensor. PLoS ONE 2013, 8, e82009. [Google Scholar] [CrossRef]
  35. Van Dongen, E.M.W.M.; Dekkers, L.M.; Spijker, K.; Meijer, E.W.; Klomp, L.W.J.; Merkx, M. Ratiometric Fluorescent Sensor Proteins with Subnanomolar Affinity for Zn(II) Based on Copper Chaperone Domains. J. Am. Chem. Soc. 2006, 128, 10754–10762. [Google Scholar] [CrossRef]
  36. Wegner, S.V.; Arslan, H.; Sunbul, M.; Yin, J.; He, C. Dynamic Copper(I) Imaging in Mammalian Cells with a Genetically Encoded Fluorescent Copper(I) Sensor. J. Am. Chem. Soc. 2010, 132, 2567–2569. [Google Scholar] [CrossRef]
  37. Lindenburg, L.; Merkx, M. Engineering genetically encoded FRET sensors. Sensors 2014, 14, 11691–11713. [Google Scholar] [CrossRef]
  38. Eroglu, E.; Gottschalk, B.; Charoensin, S.; Blass, S.; Bischof, H.; Rost, R.; Madreiter-Sokolowski, C.T.; Pelzmann, B.; Bernhart, E.; Sattler, W.; et al. Development of novel FP-based probes for live-cell imaging of nitric oxide dynamics. Nat. Comm. 2016, 7, 10623. [Google Scholar] [CrossRef]
  39. Yaginuma, H.; Kawai, S.; Tabata, K.V.; Tomiyama, K.; Kakizuka, A.; Komatsuzaki, T.; Noji, H.; Imamura, H. Diversity in ATP concentrations in a single bacterial cell population revealed by quantitative single-cell imaging. Sci. Rep. 2015, 4. [Google Scholar] [CrossRef]
  40. Shen, Y.; Rosendale, M.; Campbell, R.E.; Perrais, D. pHuji, a pH-sensitive red fluorescent protein for imaging of exo- and endocytosis. J. Cell Biol. 2014, 207, 419–432. [Google Scholar] [CrossRef] [Green Version]
  41. Wu, J.; Liu, L.; Matsuda, T.; Zhao, Y.; Rebane, A.; Drobizhev, M.; Chang, Y.-F.; Araki, S.; Arai, Y.; March, K.; et al. Improved orange and red Ca2± indicators and photophysical considerations for optogenetic applications. Acs Chem. Neurosci 2013, 4, 963–972. [Google Scholar] [CrossRef]
  42. Greenwald, E.C.; Mehta, S.; Zhang, J. Genetically Encoded Fluorescent Biosensors Illuminate the Spatiotemporal Regulation of Signaling Networks. Chem. Rev. 2018, 118, 11707–11794. [Google Scholar] [CrossRef] [PubMed]
  43. Leavesley, S.J.; Rich, T.C. Overcoming limitations of FRET measurements. Cytometry 2016, 89, 325–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zhao, Y.; Araki, S.; Wu, J.; Teramoto, T.; Chang, Y.-F.; Nakano, M.; Abdelfattah, A.S.; Fujiwara, M.; Ishihara, T.; Nagai, T.; et al. An expanded palette of genetically encoded Ca2+ indicators. Science 2011, 333, 1888–1891. [Google Scholar] [CrossRef]
  45. Pérez Koldenkova, V.; Nagai, T. Genetically encoded Ca2+ indicators: Properties and evaluation. Biochim. Et Biophys. ACTA 2013, 1833, 1787–1797. [Google Scholar] [CrossRef] [PubMed]
  46. Woehler, A.; Wlodarczyk, J.; Neher, E. Signal/Noise Analysis of FRET-Based Sensors. Biophys. J. 2010, 99, 2344–2354. [Google Scholar] [CrossRef] [Green Version]
  47. Palmer, A.E.; Qin, Y.; Park, J.G.; McCombs, J.E. Design and application of genetically encoded biosensors. Trends Biotechnol. 2011, 29, 144–152. [Google Scholar] [CrossRef] [Green Version]
  48. Waldeck-Weiermair, M.; Alam, M.R.; Khan, M.J.; Deak, A.T.; Vishnu, N.; Karsten, F.; Imamura, H.; Graier, W.F.; Malli, R. Spatiotemporal Correlations between Cytosolic and Mitochondrial Ca2+ Signals Using a Novel Red-Shifted Mitochondrial Targeted Cameleon. PLoS ONE 2012, 7, e45917. [Google Scholar] [CrossRef]
  49. Tivey, D.R.; Simmons, N.L.; Aiton, J.F. Role of passive potassium fluxes in cell volume regulation in cultured HeLa cells. J. Membr. Biol. 1985, 87, 93–105. [Google Scholar] [CrossRef]
  50. Clausen, M.J.V.; Poulsen, H. Sodium/Potassium Homeostasis in the Cell. Met. Ions Life Sci. 2013, 12, 41–67. [Google Scholar]
  51. Larkin, J.M.; Brown, M.S.; Goldstein, J.L.; Anderson, R.G. Depletion of intracellular potassium arrests coated pit formation and receptor-mediated endocytosis in fibroblasts. Cell 1983, 33, 273–285. [Google Scholar] [CrossRef]
  52. Ceccarelli, B.; Fesce, R.; Grohovaz, F.; Haimann, C. The effect of potassium on exocytosis of transmitter at the frog neuromuscular junction. J. Physiol. 1988, 401, 163–183. [Google Scholar] [CrossRef]
  53. El Kebir, D.; József, L.; Khreiss, T.; Filep, J.G. Inhibition of K+ efflux prevents mitochondrial dysfunction, and suppresses caspase-3-, apoptosis-inducing factor-, and endonuclease G-mediated constitutive apoptosis in human neutrophils. Cell. Signal. 2006, 18, 2302–2313. [Google Scholar] [CrossRef] [PubMed]
  54. Kachmar, J.F.; Boyer, P.D. Kinetic analysis of enzyme reactions. II. The potassium activation and calcium inhibition of pyruvic phosphoferase. J. Biol. Chem. 1953, 200, 669–682. [Google Scholar] [PubMed]
  55. Kato, M.; Chuang, J.L.; Tso, S.-C.; Wynn, R.M.; Chuang, D.T. Crystal structure of pyruvate dehydrogenase kinase 3 bound to lipoyl domain 2 of human pyruvate dehydrogenase complex. EMBO J. 2005, 24, 1763–1774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Shibata, N.; Masuda, J.; Tobimatsu, T.; Toraya, T.; Suto, K.; Morimoto, Y.; Yasuoka, N. A new mode of B12 binding and the direct participation of a potassium ion in enzyme catalysis: X-ray structure of diol dehydratase. Structure 1999, 7, 997–1008. [Google Scholar] [CrossRef] [Green Version]
  57. Liao, D.-I.; Dotson, G.; Turner, I.; Reiss, L.; Emptage, M. Crystal structure of substrate free form of glycerol dehydratase. J. Inorg. Biochem. 2003, 93, 84–91. [Google Scholar] [CrossRef]
  58. Andersson, C.E.; Mowbray, S.L. Activation of ribokinase by monovalent cations. J. Mol. Biol. 2002, 315, 409–419. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Dougherty, M.; Downs, D.M.; Ealick, S.E. Crystal structure of an aminoimidazole riboside kinase from Salmonella enterica: Implications for the evolution of the ribokinase superfamily. Structure 2004, 12, 1809–1821. [Google Scholar] [CrossRef]
  60. Jezek, P.; Mahdi, F.; Garlid, K.D. Reconstitution of the beef heart and rat liver mitochondrial K+/H+ (Na+/H+) antiporter. Quantitation of K+ transport with the novel fluorescent probe, PBFI. J. Biol. Chem. 1990, 265, 10522–10526. [Google Scholar]
  61. Rimmele, T.S.; Chatton, J.-Y. A Novel Optical Intracellular Imaging Approach for Potassium Dynamics in Astrocytes. PLoS ONE 2014, 9, e109243. [Google Scholar] [CrossRef]
  62. Kong, X.; Su, F.; Zhang, L.; Yaron, J.; Lee, F.; Shi, Z.; Tian, Y.; Meldrum, D.R. A Highly Selective Mitochondria-Targeting Fluorescent K+ Sensor. Angew. Chem. Int. Ed. 2015, 54, 12053–12057. [Google Scholar] [CrossRef]
  63. Ashraf, K.U.; Josts, I.; Mosbahi, K.; Kelly, S.M.; Byron, O.; Smith, B.O.; Walker, D. The Potassium Binding Protein Kbp Is a Cytoplasmic Potassium Sensor. Structure 2016, 24, 741–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Weber, A.; Jung, K. Profiling Early Osmostress-Dependent Gene Expression in Escherichia coli Using DNA Macroarrays. J. Bacteriol. 2002, 184, 5502–5507. [Google Scholar] [CrossRef]
  65. Storvik, K.A.M.; Foster, P.L. RpoS, the Stress Response Sigma Factor, Plays a Dual Role in the Regulation of Escherichia coli’s Error-Prone DNA Polymerase IV. J. Bacteriol. 2010, 192, 3639–3644. [Google Scholar] [CrossRef]
  66. Shen, Y.; Wu, S.-Y.; Rancic, V.; Aggarwal, A.; Qian, Y.; Miyashita, S.-I.; Ballanyi, K.; Campbell, R.E.; Dong, M. Genetically encoded fluorescent indicators for imaging intracellular potassium ion concentration. Commun. Biol. 2019, 2. [Google Scholar] [CrossRef]
  67. Palmer, B.F. Regulation of Potassium Homeostasis. Clin. J. Am. Soc. Nephrol. 2015, 10, 1050–1060. [Google Scholar] [CrossRef]
  68. Thier, S.O. Potassium physiology. Am. J. Med. 1986, 80, 3–7. [Google Scholar] [CrossRef]
  69. Romani, A.M.P. Cellular magnesium homeostasis. Arch. Biochem. Biophys. 2011, 512, 1–23. [Google Scholar] [CrossRef] [Green Version]
  70. Vernon, W.B. The role of magnesium in nucleic-acid and protein metabolism. Magnesium 1988, 7, 234–248. [Google Scholar]
  71. Walker, G.M. Magnesium and cell cycle control: An update. Magnesium 1986, 5, 9–23. [Google Scholar]
  72. Noronha, J.L.; Matuschak, G.M. Magnesium in critical illness: Metabolism, assessment, and treatment. Intensive Care Med. 2002, 28, 667–679. [Google Scholar] [CrossRef]
  73. Bara, M.; Guiet-Bara, A.; Durlach, J. Regulation of sodium and potassium pathways by magnesium in cell membranes. Magnes Res. 1993, 6, 167–177. [Google Scholar]
  74. Jahnen-Dechent, W.; Ketteler, M. Magnesium basics. Clin. Kidney J. 2012, 5, i3–i14. [Google Scholar] [CrossRef] [Green Version]
  75. Dugal, B.S.; Louis, B.M. Effect of magnesium ion (Mg2+) and the magnesium adenosine triphosphate ion (MgATP2-) on pigeon liver pyruvate carboxylase. Enzyme 1975, 20, 98–110. [Google Scholar] [CrossRef] [PubMed]
  76. Garfinkel, L.; Garfinkel, D. Magnesium regulation of the glycolytic pathway and the enzymes involved. Magnesium 1985, 4, 60–72. [Google Scholar] [PubMed]
  77. Stangherlin, A.; O’Neill, J.S. Signal Transduction: Magnesium Manifests as a Second Messenger. Curr. Biol. 2018, 28, R1403–R1405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Leehey, D.J. Magnesium Homeostasis in CKD. Adv. Chronic Kidney Dis. 2018, 25, 222–223. [Google Scholar] [CrossRef]
  79. Mountokalakis, T.D. Magnesium metabolism in chronic renal failure. Magnes Res. 1990, 3, 121–127. [Google Scholar] [PubMed]
  80. Bilbey, D.L.; Prabhakaran, V.M. Muscle cramps and magnesium deficiency: Case reports. Can. Fam. Physician 1996, 42, 1348–1351. [Google Scholar]
  81. Nuytten, D.; Van Hees, J.; Meulemans, A.; Carton, H. Magnesium deficiency as a cause of acute intractable seizures. J. Neurol. 1991, 238, 262–264. [Google Scholar] [CrossRef]
  82. Killilea, D.W.; Ames, B.N. Magnesium deficiency accelerates cellular senescence in cultured human fibroblasts. Proc. Natl. Acad. Sci. USA 2008, 105, 5768–5773. [Google Scholar] [CrossRef] [Green Version]
  83. Tangvoraphonkchai, K.; Davenport, A. Magnesium and Cardiovascular Disease. Adv. Chronic Kidney Dis. 2018, 25, 251–260. [Google Scholar] [CrossRef]
  84. Galassi, A.; Cozzolino, M. Magnesium: A renewed player of vascular ageing in diabetic CKD patients? Clin. Kidney J. 2014, 7, 93–96. [Google Scholar] [CrossRef] [PubMed]
  85. Raju, B.; Murphy, E.; Levy, L.A.; Hall, R.D.; London, R.E. A fluorescent indicator for measuring cytosolic free magnesium. Am. J. Physiol. Cell Physiol. 1989, 256, C540–C548. [Google Scholar] [CrossRef]
  86. Salisbury, J.L. Centrin, centrosomes, and mitotic spindle poles. Curr. Opin. Cell Biol. 1995, 7, 39–45. [Google Scholar] [CrossRef]
  87. Cox, J.A.; Tirone, F.; Durussel, I.; Firanescu, C.; Blouquit, Y.; Duchambon, P.; Craescu, C.T. Calcium and Magnesium Binding to Human Centrin 3 and Interaction with Target Peptides. Biochemistry 2005, 44, 840–850. [Google Scholar] [CrossRef] [PubMed]
  88. Maeshima, K.; Matsuda, T.; Shindo, Y.; Imamura, H.; Tamura, S.; Imai, R.; Kawakami, S.; Nagashima, R.; Soga, T.; Noji, H.; et al. A Transient Rise in Free Mg2+ Ions Released from ATP-Mg Hydrolysis Contributes to Mitotic Chromosome Condensation. Curr. Biol. 2018, 28, 444–451. [Google Scholar] [CrossRef]
  89. Lunin, V.V.; Dobrovetsky, E.; Khutoreskaya, G.; Zhang, R.; Joachimiak, A.; Doyle, D.A.; Bochkarev, A.; Maguire, M.E.; Edwards, A.M.; Koth, C.M. Crystal structure of the CorA Mg2+ transporter. Nature 2006, 440, 833–837. [Google Scholar] [CrossRef]
  90. Smith, R.L.; Banks, J.L.; Snavely, M.D.; Maguire, M.E. Sequence and topology of the CorA magnesium transport systems of Salmonella typhimurium and Escherichia coli. Identification of a new class of transport protein. J. Biol. Chem. 1993, 268, 14071–14080. [Google Scholar] [PubMed]
  91. Sheu, S.S.; Sharma, V.K.; Banerjee, S.P. Measurement of cytosolic free calcium concentration in isolated rat ventricular myocytes with quin 2. Circ. Res. 1984, 55, 830–834. [Google Scholar] [CrossRef] [PubMed]
  92. Clapham, D.E. Calcium Signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef] [Green Version]
  93. Ahvazi, B.; Boeshans, K.M.; Idler, W.; Baxa, U.; Steinert, P.M. Roles of Calcium Ions in the Activation and Activity of the Transglutaminase 3 Enzyme. J. Biol. Chem. 2003, 278, 23834–23841. [Google Scholar] [CrossRef] [Green Version]
  94. Lee, S.; Park, H.I.; Sang, Q.-X.A. Calcium regulates tertiary structure and enzymatic activity of human endometase/matrilysin-2 and its role in promoting human breast cancer cell invasion. Biochem. J. 2007, 403, 31–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Shah, V.N.; Chagot, B.; Chazin, W.J. Calcium-Dependent Regulation of Ion Channels. Calcium Bind Proteins 2006, 1, 203–212. [Google Scholar] [PubMed]
  96. Hedrich, R.; Neher, E. Cytoplasmic calcium regulates voltage-dependent ion channels in plant vacuoles. Nature 1987, 329, 833–836. [Google Scholar] [CrossRef]
  97. Van Haasteren, G.; Li, S.; Muda, M.; Susini, S.; Schlegel, W. Calcium Signalling and Gene Expression. J. Recept. Signal. Transduct. 1999, 19, 481–492. [Google Scholar] [CrossRef]
  98. Kuo, I.Y.; Ehrlich, B.E. Signaling in muscle contraction. Cold Spring Harb. Perspect. Biol. 2015, 7, a006023. [Google Scholar] [CrossRef]
  99. Südhof, T.C. Calcium control of neurotransmitter release. Cold Spring Harb. Perspect. Biol. 2012, 4, a011353. [Google Scholar] [CrossRef]
  100. Malaisse, W.J. Role of Calcium in the Regulation of Hormonal Secretion. Horm. Res. 1984, 20, 28–37. [Google Scholar] [CrossRef]
  101. Whitaker, M. Calcium at Fertilization and in Early Development. Physiol. Rev. 2006, 86, 25–88. [Google Scholar] [CrossRef] [PubMed]
  102. Giorgi, C.; Marchi, S.; Pinton, P. The machineries, regulation and cellular functions of mitochondrial calcium. Nat. Rev. Mol.ol. Cell Biol. 2018, 19, 713–730. [Google Scholar] [CrossRef]
  103. Treiman, M. Regulation of the endoplasmic reticulum calcium storage during the unfolded protein response--significance in tissue ischemia? Trends Cardiovasc. Med. 2002, 12, 57–62. [Google Scholar] [CrossRef]
  104. Bustos, G.; Cruz, P.; Lovy, A.; Cárdenas, C. Endoplasmic Reticulum–Mitochondria Calcium Communication and the Regulation of Mitochondrial Metabolism in Cancer: A Novel Potential Target. Front. Oncol. 2017, 7, 199. [Google Scholar] [CrossRef]
  105. Prins, D.; Michalak, M. Organellar Calcium Buffers. Cold Spring Harb. Perspect. Biol. 2011, 3, a004069. [Google Scholar] [CrossRef]
  106. Koch, G.L. The endoplasmic reticulum and calcium storage. Bioessays 1990, 12, 527–531. [Google Scholar] [CrossRef] [PubMed]
  107. Coe, H.; Michalak, M. Calcium binding chaperones of the endoplasmic reticulum. Gen. Physiol. Biophys. 2009, 28, F96–F103. [Google Scholar] [PubMed]
  108. Pedriali, G.; Rimessi, A.; Sbano, L.; Giorgi, C.; Wieckowski, M.R.; Previati, M.; Pinton, P. Regulation of Endoplasmic Reticulum–Mitochondria Ca2+ Transfer and Its Importance for Anti-Cancer Therapies. Front. Oncol. 2017, 7, 180. [Google Scholar] [CrossRef]
  109. Duchen, M.R. Mitochondria and calcium: From cell signalling to cell death. J. Physiol. 2000, 529, 57–68. [Google Scholar] [CrossRef]
  110. Giorgi, C.; Baldassari, F.; Bononi, A.; Bonora, M.; De Marchi, E.; Marchi, S.; Missiroli, S.; Patergnani, S.; Rimessi, A.; Suski, J.M.; et al. Mitochondrial Ca2+ and apoptosis. Cell Calcium 2012, 52, 36–43. [Google Scholar] [CrossRef]
  111. Santo-Domingo, J.; Demaurex, N. The renaissance of mitochondrial pH. J. Gen. Physiol. 2012, 139, 415–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Graier, W.F.; Frieden, M.; Malli, R. Mitochondria and Ca2+ signaling: Old guests, new functions. Pflug. Arch. Eur J. Physiol. 2007, 455, 375–396. [Google Scholar] [CrossRef] [PubMed]
  113. Williams, G.S.B.; Boyman, L.; Chikando, A.C.; Khairallah, R.J.; Lederer, W.J. Mitochondrial calcium uptake. Proc. Natl. Acad. Sci. USA 2013, 110, 10479–10486. [Google Scholar] [CrossRef] [Green Version]
  114. Grynkiewicz, G.; Poenie, M.; Tsien, R.Y. A new generation of Ca2+ indicators with greatly improved fluorescence properties. J. Biol. Chem. 1985, 260, 3440–3450. [Google Scholar]
  115. Griesbeck, O.; Baird, G.S.; Campbell, R.E.; Zacharias, D.A.; Tsien, R.Y. Reducing the Environmental Sensitivity of Yellow Fluorescent Protein: MECHANISM AND APPLICATIONS. J. Biol. Chem. 2001, 276, 29188–29194. [Google Scholar] [CrossRef] [PubMed]
  116. Truong, K.; Sawano, A.; Mizuno, H.; Hama, H.; Tong, K.I.; Mal, T.K.; Miyawaki, A.; Ikura, M. FRET-based in vivo Ca2+ imaging by a new calmodulin-GFP fusion molecule. Nat. Struct. Biol. 2001, 8, 1069–1073. [Google Scholar] [CrossRef]
  117. Nagai, T.; Yamada, S.; Tominaga, T.; Ichikawa, M.; Miyawaki, A. Expanded dynamic range of fluorescent indicators for Ca2+ by circularly permuted yellow fluorescent proteins. Proc. Natl. Acad. Sci. USA 2004, 101, 10554–10559. [Google Scholar] [CrossRef]
  118. Palmer, A.E.; Jin, C.; Reed, J.C.; Tsien, R.Y. Bcl-2-mediated alterations in endoplasmic reticulum Ca2+ analyzed with an improved genetically encoded fluorescent sensor. Proc. Natl. Acad. Sci. USA 2004, 101, 17404–17409. [Google Scholar] [CrossRef] [PubMed]
  119. Palmer, A.E.; Giacomello, M.; Kortemme, T.; Hires, S.A.; Lev-Ram, V.; Baker, D.; Tsien, R.Y. Ca2+ Indicators Based on Computationally Redesigned Calmodulin-Peptide Pairs. Chem. Biol. 2006, 13, 521–530. [Google Scholar] [CrossRef]
  120. Giacomello, M.; Drago, I.; Bortolozzi, M.; Scorzeto, M.; Gianelle, A.; Pizzo, P.; Pozzan, T. Ca2+ Hot Spots on the Mitochondrial Surface Are Generated by Ca2+ Mobilization from Stores, but Not by Activation of Store-Operated Ca2+ Channels. Mol. Cell 2010, 38, 280–290. [Google Scholar] [CrossRef]
  121. Ishii, K.; Hirose, K.; Iino, M. Ca2+ shuttling between endoplasmic reticulum and mitochondria underlying Ca2+ oscillations. EMBO Rep. 2006, 7, 390–396. [Google Scholar] [CrossRef]
  122. Tay, L.H.; Dick, I.E.; Yang, W.; Mank, M.; Griesbeck, O.; Yue, D.T. Nanodomain Ca2+ of Ca2+ channels detected by a tethered genetically encoded Ca2+ sensor. Nat. Commun. 2012, 3, 778. [Google Scholar] [CrossRef]
  123. Heim, N.; Griesbeck, O. Genetically Encoded Indicators of Cellular Calcium Dynamics Based on Troponin C and Green Fluorescent Protein. J. Biol. Chem. 2004, 279, 14280–14286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Mank, M.; Reiff, D.F.; Heim, N.; Friedrich, M.W.; Borst, A.; Griesbeck, O. A FRET-Based Calcium Biosensor with Fast Signal Kinetics and High Fluorescence Change. Biophys. J. 2006, 90, 1790–1796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Herzberg, O.; Moult, J.; James, M.N. Calcium binding to skeletal muscle troponin C and the regulation of muscle contraction. CIBA Found. Symp. 1986, 122, 120–144. [Google Scholar]
  126. Osibow, K.; Malli, R.; Kostner, G.M.; Graier, W.F. A new type of non-Ca2+-buffering Apo(a)-based fluorescent indicator for intraluminal Ca2+ in the endoplasmic reticulum. J. Biol. Chem. 2006, 281, 5017–5025. [Google Scholar] [CrossRef]
  127. Nagai, T.; Sawano, A.; Park, E.S.; Miyawaki, A. Circularly permuted green fluorescent proteins engineered to sense Ca2+. Proc. Natl. Acad. Sci. USA 2001, 98, 3197–3202. [Google Scholar] [CrossRef]
  128. Suzuki, J.; Kanemaru, K.; Iino, M. Genetically Encoded Fluorescent Indicators for Organellar Calcium Imaging. Biophys. J. 2016, 111, 1119–1131. [Google Scholar] [CrossRef] [Green Version]
  129. Miyawaki, A.; Griesbeck, O.; Heim, R.; Tsien, R.Y. Dynamic and quantitative Ca2+ measurements using improved cameleons. Proc. Natl. Acad. Sci. USA 1999, 96, 2135–2140. [Google Scholar] [CrossRef]
  130. Hordyjewska, A.; Popiołek, Ł.; Kocot, J. The many “faces” of copper in medicine and treatment. Biometals 2014, 27, 611–621. [Google Scholar] [CrossRef] [PubMed]
  131. Fatemi, N.; Sarkar, B. Molecular mechanism of copper transport in Wilson disease. Environ. Health Perspect. 2002, 110, 695–698. [Google Scholar] [CrossRef]
  132. Tümer, Z.; Møller, L.B. Menkes disease. Eur J. Hum. Genet. 2010, 18, 511–518. [Google Scholar] [CrossRef]
  133. Pham, A.N.; Xing, G.; Miller, C.J.; Waite, T.D. Fenton-like copper redox chemistry revisited: Hydrogen peroxide and superoxide mediation of copper-catalyzed oxidant production. J. Catal. 2013, 301, 54–64. [Google Scholar] [CrossRef]
  134. Gaetke, L.M.; Chow-Johnson, H.S.; Chow, C.K. Copper: Toxicological relevance and mechanisms. Arch. Toxicol. 2014, 88, 1929–1938. [Google Scholar] [CrossRef]
  135. Klevay, L.M. Cardiovascular Disease from Copper Deficiency—A History. J. Nutr. 2000, 130, 489S–492S. [Google Scholar] [CrossRef]
  136. Waggoner, D.J.; Bartnikas, T.B.; Gitlin, J.D. The role of copper in neurodegenerative disease. Neurobiol. Dis. 1999, 6, 221–230. [Google Scholar] [CrossRef] [PubMed]
  137. Inesi, G. Molecular features of copper binding proteins involved in copper homeostasis: Copper Binding Proteins. IUBMB Life 2017, 69, 211–217. [Google Scholar] [CrossRef] [PubMed]
  138. Pang, W.L.; Kaur, A.; Ratushny, A.V.; Cvetkovic, A.; Kumar, S.; Pan, M.; Arkin, A.P.; Aitchison, J.D.; Adams, M.W.W.; Baliga, N.S. Metallochaperones Regulate Intracellular Copper Levels. PLoS Comput. Biol. 2013, 9, e1002880. [Google Scholar] [CrossRef]
  139. Waldron, K.J.; Robinson, N.J. How do bacterial cells ensure that metalloproteins get the correct metal? Nat. Rev. Microbiol. 2009, 7, 25–35. [Google Scholar] [CrossRef]
  140. Carter, K.P.; Young, A.M.; Palmer, A.E. Fluorescent Sensors for Measuring Metal Ions in Living Systems. Chem. Rev. 2014, 114, 4564–4601. [Google Scholar] [CrossRef] [PubMed]
  141. Koch, K.A.; Allard, S.; Santoro, N.; Côté, J.; Thiele, D.J. The Candida glabrata Amt1 copper-sensing transcription factor requires Swi/Snf and Gcn5 at a critical step in copper detoxification. Mol. Microbiol. 2001, 40, 1165–1174. [Google Scholar] [CrossRef]
  142. Wegner, S.V.; Sun, F.; Hernandez, N.; He, C. The tightly regulated copper window in yeast. Chem. Commun. 2011, 47, 2571–2573. [Google Scholar] [CrossRef]
  143. Gralla, E.B.; Thiele, D.J.; Silar, P.; Valentine, J.S. ACE1, a copper-dependent transcription factor, activates expression of the yeast copper, zinc superoxide dismutase gene. Proc. Natl. Acad. Sci. USA 1991, 88, 8558–8562. [Google Scholar] [CrossRef] [PubMed]
  144. Keller, G.; Bird, A.; Winge, D.R. Independent Metalloregulation of Ace1 and Mac1 in Saccharomyces cerevisiae. Eukaryot. Cell 2005, 4, 1863–1871. [Google Scholar] [CrossRef]
  145. Zhu, Z.; Labbé, S.; Peña, M.M.O.; Thiele, D.J. Copper Differentially Regulates the Activity and Degradation of Yeast Mac1 Transcription Factor. J. Biol. Chem. 1998, 273, 1277–1280. [Google Scholar] [CrossRef] [Green Version]
  146. Yan, X.; Li, X.; Lv, S.-S.; He, D.-C. A novel genetically encoded fluorescent protein as a Cu(I) indicator. Dalton Trans. 2012, 41, 727–729. [Google Scholar] [CrossRef] [PubMed]
  147. Festa, R.A.; Jones, M.B.; Butler-Wu, S.; Sinsimer, D.; Gerads, R.; Bishai, W.R.; Peterson, S.N.; Darwin, K.H. A novel copper-responsive regulon in Mycobacterium tuberculosis: RicR regulon in Mycobacterium tuberculosis. Mol. Microbiol. 2011, 79, 133–148. [Google Scholar] [CrossRef]
  148. Ravimohan, S.; Kornfeld, H.; Weissman, D.; Bisson, G.P. Tuberculosis and lung damage: From epidemiology to pathophysiology. Eur. Respir. Rev. 2018, 27, 170077. [Google Scholar] [CrossRef]
  149. Lim, Y.S.; Kim, S.B.; Kim, M.K.; Lim, Y.J. Disseminated Tuberculosis of Central Nervous System: Spinal Intramedullary and Intracranial Tuberculomas. J. Korean Neurosurg Soc. 2013, 54, 61. [Google Scholar] [CrossRef]
  150. Malaviya, A.N.; Kotwal, P.P. Arthritis associated with tuberculosis. Best Pr. Res. Clin. Rheumatol. 2003, 17, 319–343. [Google Scholar] [CrossRef]
  151. Daher, E.D.F.; Barros, E.J.G.; da Silva, G.B., Jr. Renal Tuberculosis in the Modern Era. Am. J. Trop. Med. Hyg. 2013, 88, 54–64. [Google Scholar] [CrossRef] [Green Version]
  152. Bishai, W.R.; Dannenberg, A.M.; Parrish, N.; Ruiz, R.; Chen, P.; Zook, B.C.; Johnson, W.; Boles, J.W.; Pitt, M.L. Virulence of Mycobacterium tuberculosis CDC1551 and H37Rv in rabbits evaluated by Lurie’s pulmonary tubercle count method. Infect. Immun. 1999, 67, 4931–4934. [Google Scholar] [PubMed]
  153. Koay, M.S.; Janssen, B.M.G.; Merkx, M. Tuning the metal binding site specificity of a fluorescent sensor protein: From copper to zinc and back. Dalton Trans. 2013, 42, 3230–3232. [Google Scholar] [CrossRef] [PubMed]
  154. Vallee, B.L.; Coleman, J.E.; Auld, D.S. Zinc fingers, zinc clusters, and zinc twists in DNA-binding protein domains. Proc. Natl. Acad. Sci. USA 1991, 88, 999–1003. [Google Scholar] [CrossRef] [PubMed]
  155. Wu, X.-F. Non-Redox-Metal-Catalyzed Redox Reactions: Zinc Catalysts. Chem. Asian J. 2012, 7, 2502–2509. [Google Scholar] [CrossRef]
  156. Tóth, K. Zinc in neurotransmission. Annu. Rev. Nutr. 2011, 31, 139–153. [Google Scholar] [CrossRef]
  157. Andreini, C.; Banci, L.; Bertini, I.; Rosato, A. Counting the Zinc-Proteins Encoded in the Human Genome. J. Proteome Res. 2006, 5, 196–201. [Google Scholar] [CrossRef] [PubMed]
  158. Borovanský, J.; Riley, P.A. Cytotoxicity of zinc in vitro. Chem. -Biol. Interact. 1989, 69, 279–291. [Google Scholar]
  159. Vinkenborg, J.L.; Nicolson, T.J.; Bellomo, E.A.; Koay, M.S.; Rutter, G.A.; Merkx, M. Genetically encoded FRET sensors to monitor intracellular Zn2+ homeostasis. Nat. Methods 2009, 6, 737–740. [Google Scholar] [CrossRef]
  160. Hatori, Y.; Lutsenko, S. The Role of Copper Chaperone Atox1 in Coupling Redox Homeostasis to Intracellular Copper Distribution. Antioxidants 2016, 5, 25. [Google Scholar] [CrossRef]
  161. Bartee, M.Y.; Lutsenko, S. Hepatic copper-transporting ATPase ATP7B: Function and inactivation at the molecular and cellular level. Biometals 2007, 20, 627–637. [Google Scholar] [CrossRef]
  162. Van Dongen, E.M.W.M.; Evers, T.H.; Dekkers, L.M.; Meijer, E.W.; Klomp, L.W.J.; Merkx, M. Variation of Linker Length in Ratiometric Fluorescent Sensor Proteins Allows Rational Tuning of Zn(II) Affinity in the Picomolar to Femtomolar Range. J. Am. Chem. Soc. 2007, 129, 3494–3495. [Google Scholar] [CrossRef]
  163. Dittmer, P.J.; Miranda, J.G.; Gorski, J.A.; Palmer, A.E. Genetically Encoded Sensors to Elucidate Spatial Distribution of Cellular Zinc. J. Biol. Chem. 2009, 284, 16289–16297. [Google Scholar] [CrossRef] [Green Version]
  164. Elrod-Erickson, M.; Rould, M.A.; Nekludova, L.; Pabo, C.O. Zif268 protein–DNA complex refined at 1.6å: A model system for understanding zinc finger–DNA interactions. Structure 1996, 4, 1171–1180. [Google Scholar] [CrossRef]
  165. Razin, S.V.; Borunova, V.V.; Maksimenko, O.G.; Kantidze, O.L. Cys2His2 zinc finger protein family: Classification, functions, and major members. Biochem. Mosc. 2012, 77, 217–226. [Google Scholar] [CrossRef] [PubMed]
  166. Matthews, J.M.; Kowalski, K.; Liew, C.K.; Sharpe, B.K.; Fox, A.H.; Crossley, M.; MacKay, J.P. A class of zinc fingers involved in protein-protein interactions biophysical characterization of CCHC fingers from fog and U-shaped. Eur. J. Biochem. 2000, 267, 1030–1038. [Google Scholar] [CrossRef] [PubMed]
  167. Hall, T.M.T. Multiple modes of RNA recognition by zinc finger proteins. Curr. Opin. Struct. Biol. 2005, 15, 367–373. [Google Scholar] [CrossRef] [Green Version]
  168. Evers, T.H.; Appelhof, M.A.M.; de Graaf-Heuvelmans, P.T.H.M.; Meijer, E.W.; Merkx, M. Ratiometric Detection of Zn(II) Using Chelating Fluorescent Protein Chimeras. J. Mol. Biol. 2007, 374, 411–425. [Google Scholar] [CrossRef]
  169. Hessels, A.M.; Chabosseau, P.; Bakker, M.H.; Engelen, W.; Rutter, G.A.; Taylor, K.M.; Merkx, M. eZinCh-2: A Versatile, Genetically Encoded FRET Sensor for Cytosolic and Intraorganelle Zn2+ Imaging. ACS Chem. Biol. 2015, 10, 2126–2134. [Google Scholar] [CrossRef]
  170. Qin, Y.; Dittmer, P.J.; Park, J.G.; Jansen, K.B.; Palmer, A.E. Measuring steady-state and dynamic endoplasmic reticulum and Golgi Zn2+ with genetically encoded sensors. Proc. Natl. Acad. Sci. USA 2011, 108, 7351–7356. [Google Scholar] [CrossRef]
  171. Miranda, J.G.; Weaver, A.L.; Qin, Y.; Park, J.G.; Stoddard, C.I.; Lin, M.Z.; Palmer, A.E. New Alternately Colored FRET Sensors for Simultaneous Monitoring of Zn2+ in Multiple Cellular Locations. PLoS ONE 2012, 7, e49371. [Google Scholar] [CrossRef]
  172. Berret, E.; Smith, P.Y.; Henry, M.; Soulet, D.; Hébert, S.S.; Toth, K.; Mouginot, D.; Drolet, G. Extracellular Na+ levels regulate formation and activity of the NaX/alpha1-Na+/K+-ATPase complex in neuronal cells. Front. Cell Neurosci. 2014, 8, 413. [Google Scholar] [CrossRef]
  173. Graf, J.; Petersen, O.H. Cell membrane potential and resistance in liver. J. Physiol. 1978, 284, 105–126. [Google Scholar] [CrossRef]
  174. De Hert, S.; De Baerdemaeker, L.; De Maeseneer, M. What the phlebologist should know about local anesthetics. Phlebology 2014, 29, 428–441. [Google Scholar] [CrossRef] [PubMed]
  175. Blaustein, M.P. Sodium ions, calcium ions, blood pressure regulation, and hypertension: A reassessment and a hypothesis. Am. J. Physiol. Cell Physiol. 1977, 232, C165–C173. [Google Scholar] [CrossRef]
  176. Huntington, J.A. How Na+ activates thrombin–a review of the functional and structural data. Biol. Chem. 2008, 389, 1025–1035. [Google Scholar] [CrossRef]
  177. Minta, A.; Tsien, R.Y. Fluorescent indicators for cytosolic sodium. J. Biol. Chem. 1989, 264, 19449–19457. [Google Scholar] [PubMed]
  178. Szmacinski, H.; Lakowicz, J.R. Sodium Green as a Potential Probe for Intracellular Sodium Imaging Based on Fluorescence Lifetime. Anal. Biochem. 1997, 250, 131–138. [Google Scholar] [CrossRef]
  179. Meier, S.D.; Kovalchuk, Y.; Rose, C.R. Properties of the new fluorescent Na+ indicator CoroNa Green: Comparison with SBFI and confocal Na+ imaging. J. Neurosci. Methods 2006, 155, 251–259. [Google Scholar] [CrossRef]
  180. Kim, M.K.; Lim, C.S.; Hong, J.T.; Han, J.H.; Jang, H.-Y.; Kim, H.M.; Cho, B.R. Sodium-Ion-Selective Two-Photon Fluorescent Probe for In Vivo Imaging. Angew. Chem. Int. Ed. 2010, 49, 364–367. [Google Scholar] [CrossRef]
  181. Sarkar, A.R.; Heo, C.H.; Park, M.Y.; Lee, H.W.; Kim, H.M. A small molecule two-photon fluorescent probe for intracellular sodium ions. Chem. Commun. 2014, 50, 1309–1312. [Google Scholar] [CrossRef]
  182. Baron, S.; Caplanusi, A.; van de Ven, M.; Radu, M.; Despa, S.; Lambrichts, I.; Ameloot, M.; Steels, P.; Smets, I. Role of Mitochondrial Na+ Concentration, Measured by CoroNa Red, in the Protection of Metabolically Inhibited MDCK Cells. JASN 2005, 16, 3490–3497. [Google Scholar] [CrossRef]
  183. Wieland, T.; Faulstich, H. Amatoxins, phallotoxins, phallolysin, and antamanide: The biologically active components of poisonous Amanita mushrooms. Crc. Crit. Rev. Biochem. 1978, 5, 185–260. [Google Scholar] [CrossRef] [PubMed]
  184. Pineda, A.O.; Carrell, C.J.; Bush, L.A.; Prasad, S.; Caccia, S.; Chen, Z.-W.; Mathews, F.S.; Di Cera, E. Molecular Dissection of Na+ Binding to Thrombin. J. Biol. Chem. 2004, 279, 31842–31853. [Google Scholar] [CrossRef]
  185. Juers, D.H.; Rob, B.; Dugdale, M.L.; Rahimzadeh, N.; Giang, C.; Lee, M.; Matthews, B.W.; Huber, R.E. Direct and indirect roles of His-418 in metal binding and in the activity of β-galactosidase (E. coli). Protein Sci. 2009, 18, 1281–1292. [Google Scholar] [CrossRef] [PubMed]
  186. Lasocki, S.; Gaillard, T.; Rineau, E. Iron is essential for living! Crit. Care 2014, 18, 678. [Google Scholar] [CrossRef] [Green Version]
  187. Williams, R.J. Free manganese (II) and iron (II) cations can act as intracellular cell controls. FEBS Lett. 1982, 140, 3–10. [Google Scholar] [CrossRef]
  188. Nappi, A.J.; Vass, E. Interactions of Iron with Reactive Intermediates of Oxygen and Nitrogen. Dev. Neurosci 2002, 24, 134–142. [Google Scholar] [CrossRef] [PubMed]
  189. Jensen, K.P.; Ryde, U. How O2 Binds to Heme: Reasons for rapid binding and spin inversion. J. Biol. Chem. 2004, 279, 14561–14569. [Google Scholar] [CrossRef]
  190. Gilardi, G.; Di Nardo, G. Heme iron centers in cytochrome P450: Structure and catalytic activity. Rend. Fis. Acc. Lincei 2017, 28, 159–167. [Google Scholar] [CrossRef]
  191. Liu, J.; Chakraborty, S.; Hosseinzadeh, P.; Yu, Y.; Tian, S.; Petrik, I.; Bhagi, A.; Lu, Y. Metalloproteins containing cytochrome, iron-sulfur, or copper redox centers. Chem. Rev. 2014, 114, 4366–4469. [Google Scholar] [CrossRef] [PubMed]
  192. Rubio, L.M.; Ludden, P.W. Biosynthesis of the Iron-Molybdenum Cofactor of Nitrogenase. Annu. Rev. Microbiol. 2008, 62, 93–111. [Google Scholar] [CrossRef] [Green Version]
  193. Schick, M.; Xie, X.; Ataka, K.; Kahnt, J.; Linne, U.; Shima, S. Biosynthesis of the iron-guanylylpyridinol cofactor of [Fe]-hydrogenase in methanogenic archaea as elucidated by stable-isotope labeling. J. Am. Chem. Soc. 2012, 134, 3271–3280. [Google Scholar] [CrossRef]
  194. Sanvisens, N.; Bañó, M.C.; Huang, M.; Puig, S. Regulation of Ribonucleotide Reductase in Response to Iron Deficiency. Mol. Cell 2011, 44, 759–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Hara, A.; Sawada, H.; Kato, T.; Nakayama, T.; Yamamoto, H.; Matsumoto, Y. Purification and characterization of a purple acid phosphatase from rat spleen. J. Biochem. 1984, 95, 67–74. [Google Scholar] [CrossRef] [PubMed]
  196. Spanner, M.; Weber, K.; Lanske, B.; Ihbe, A.; Siggelkow, H.; Schütze, H.; Atkinson, M.J. The iron-binding protein ferritin is expressed in cells of the osteoblastic lineage in vitro and in vivo. Bone 1995, 17, 161–165. [Google Scholar] [CrossRef]
  197. Koppenol, W.H. The centennial of the Fenton reaction. Free Radic. Biol. Med. 1993, 15, 645–651. [Google Scholar] [CrossRef]
  198. Dev, S.; Babitt, J.L. Overview of iron metabolism in health and disease: Iron metabolism in health and disease. Hemodial. Int. 2017, 21, S6–S20. [Google Scholar] [CrossRef]
  199. García-Beltrán, O.; Mena, N.; Yañez, O.; Caballero, J.; Vargas, V.; Nuñez, M.T.; Cassels, B.K. Design, synthesis and cellular dynamics studies in membranes of a new coumarin-based “turn-off” fluorescent probe selective for Fe2+. Eur. J. Med. Chem. 2013, 67, 60–63. [Google Scholar] [CrossRef]
  200. Wang, R.; Yu, F.; Liu, P.; Chen, L. A turn-on fluorescent probe based on hydroxylamine oxidation for detecting ferric ion selectively in living cells. Chem. Commun. 2012, 48, 5310. [Google Scholar] [CrossRef]
  201. Breuer, W.; Epsztejn, S.; Millgram, P.; Cabantchik, I.Z. Transport of iron and other transition metals into cells as revealed by a fluorescent probe. Am. J. Physiol. Cell Physiol. 1995, 268, C1354–C1361. [Google Scholar] [CrossRef] [PubMed]
  202. Goel, A.; Umar, S.; Nag, P.; Sharma, A.; Kumar, L.; Shamsuzzama, S.; Hossain, Z.; Gayen, J.R.; Nazir, A. A dual colorimetric-ratiometric fluorescent probe NAP-3 for selective detection and imaging of endogenous labile iron(III) pools in C. elegans. Chem. Commun. 2015, 51, 5001–5004. [Google Scholar] [CrossRef] [PubMed]
  203. Li, P.; Fang, L.; Zhou, H.; Zhang, W.; Wang, X.; Li, N.; Zhong, H.; Tang, B. A New Ratiometric Fluorescent Probe for Detection of Fe2+ with High Sensitivity and Its Intracellular Imaging Applications. Chem. Eur. J. 2011, 17, 10520–10523. [Google Scholar] [CrossRef]
  204. Kehl-Fie, T.E.; Skaar, E.P. Nutritional immunity beyond iron: A role for manganese and zinc. Curr. Opin. Chem. Biol. 2010, 14, 218–224. [Google Scholar] [CrossRef]
  205. Foster, A.W.; Osman, D.; Robinson, N.J. Metal Preferences and Metallation. J. Biol. Chem. 2014, 289, 28095–28103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Smith, M.R.; Fernandes, J.; Go, Y.-M.; Jones, D.P. Redox dynamics of manganese as a mitochondrial life-death switch. Biochem. Biophys. Res. Commun. 2017, 482, 388–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Aguirre, J.D.; Culotta, V.C. Battles with Iron: Manganese in Oxidative Stress Protection. J. Biol. Chem. 2012, 287, 13541–13548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Mukhopadhyay, S.; Bachert, C.; Smith, D.R.; Linstedt, A.D. Manganese-induced Trafficking and Turnover of the cis -Golgi Glycoprotein GPP130. MBoC 2010, 21, 1282–1292. [Google Scholar] [CrossRef] [PubMed]
  209. Takeda, A. Manganese action in brain function. Brain Res. Brain Res. Rev. 2003, 41, 79–87. [Google Scholar] [CrossRef]
  210. Crossgrove, J.; Zheng, W. Manganese toxicity upon overexposure. NMR Biomed. 2004, 17, 544–553. [Google Scholar] [CrossRef] [PubMed]
  211. Liang, J.; Canary, J.W. Discrimination between Hard Metals with Soft Ligand Donor Atoms: An On-Fluorescence Probe for Manganese(II). Angew. Chem. Int. Ed. 2010, 49, 7710–7713. [Google Scholar] [CrossRef]
  212. Guth, J.H.; Burris, R.H. The role of Mg2+ and Mn2+ in the enzyme-catalysed activation of nitrogenase Fe protein from Rhodospirillum rubrum. Biochem. J. 1983, 213, 741–749. [Google Scholar] [CrossRef]
  213. Kehres, D.G.; Maguire, M.E. Emerging themes in manganese transport, biochemistry and pathogenesis in bacteria. FEMS Microbiol. Rev. 2003, 27, 263–290. [Google Scholar] [CrossRef] [Green Version]
  214. Ceccarelli, C.; Grodsky, N.B.; Ariyaratne, N.; Colman, R.F.; Bahnson, B.J. Crystal Structure of Porcine Mitochondrial NADP+ -dependent Isocitrate Dehydrogenase Complexed with Mn2+ and Isocitrate: INSIGHTS INTO THE ENZYME MECHANISM. J. Biol. Chem. 2002, 277, 43454–43462. [Google Scholar] [CrossRef]
  215. Romero, N.; Benítez, J.; Garcia, D.; González, A.; Bennun, L.; García-Robles, M.A.; López, V.; Wilson, L.A.; Schenk, G.; Carvajal, N.; et al. Mammalian agmatinases constitute unusual members in the family of Mn2+ -dependent ureahydrolases. J. Inorg. Biochem. 2017, 166, 122–125. [Google Scholar] [CrossRef] [PubMed]
  216. Thestrup, T.; Litzlbauer, J.; Bartholomäus, I.; Mues, M.; Russo, L.; Dana, H.; Kovalchuk, Y.; Liang, Y.; Kalamakis, G.; Laukat, Y.; et al. Optimized ratiometric calcium sensors for functional in vivo imaging of neurons and T lymphocytes. Nat. Methods 2014, 11, 175–182. [Google Scholar] [CrossRef] [PubMed]
  217. Craggs, T.D. Green fluorescent protein: Structure, folding and chromophore maturation. Chem. Soc. Rev. 2009, 38, 2865–2875. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of extra- and intracellular metal ion distributions. Concentrations of the physiologically most relevant metal ions in the extracellular fluid (upper left, light blue) and various cellular organelles including the cytosol (upper right, light brown), nucleus (light yellow), endoplasmic reticulum (brown), Golgi apparatus (red), mitochondrial matrix (blue) and lysosomes (pink) are demonstrated.
Figure 1. Schematic representation of extra- and intracellular metal ion distributions. Concentrations of the physiologically most relevant metal ions in the extracellular fluid (upper left, light blue) and various cellular organelles including the cytosol (upper right, light brown), nucleus (light yellow), endoplasmic reticulum (brown), Golgi apparatus (red), mitochondrial matrix (blue) and lysosomes (pink) are demonstrated.
Cells 08 00492 g001
Figure 2. Schematic representation of the functional principle of genetically encoded probes (GEPs). Schematic representation of FRET-based indicators either containing (a) a CFP (cyan) and YFP (yellow) or (b) a GFP (green) and RFP (orange). In the ion (i, grey circles, a and b) unbound conformation FRET efficiency is low, donor fluorescence is high. Upon ion binding to the ion sensitive domain, the probes undergo a conformational rearrangement, increasing FRET-fluorescence, while the donor fluorescence, i.e., CFP or GFP, is decreasing (a and b, right panels). (c,d) schematically demonstrate the course of FRET- (yellow and orange curve, c and d), donor fluorescence (CFP, cyan and GFP, green curve, c and d), as well as the course of the resulting FRET-ratio signal of dividing FRET- by donor fluorescence (Ratio, black curves, c and d).
Figure 2. Schematic representation of the functional principle of genetically encoded probes (GEPs). Schematic representation of FRET-based indicators either containing (a) a CFP (cyan) and YFP (yellow) or (b) a GFP (green) and RFP (orange). In the ion (i, grey circles, a and b) unbound conformation FRET efficiency is low, donor fluorescence is high. Upon ion binding to the ion sensitive domain, the probes undergo a conformational rearrangement, increasing FRET-fluorescence, while the donor fluorescence, i.e., CFP or GFP, is decreasing (a and b, right panels). (c,d) schematically demonstrate the course of FRET- (yellow and orange curve, c and d), donor fluorescence (CFP, cyan and GFP, green curve, c and d), as well as the course of the resulting FRET-ratio signal of dividing FRET- by donor fluorescence (Ratio, black curves, c and d).
Cells 08 00492 g002
Figure 3. Pseudo-colored ratio images over time of HeLa cells expressing GEPII 1.0, a FRET-based K+ sensor. Pseudo-colored ratio images of HeLa cells expressing NES-GEPII 1.0 were treated with 15 µM gramicidin, a K+ selective channel forming peptide, in the absence or presence of 140 mM extracellular K+ as indicated. White numbers in the images represent the time in min:sec of image acquisition. Scale bar in the right lower image represents 10 µM.
Figure 3. Pseudo-colored ratio images over time of HeLa cells expressing GEPII 1.0, a FRET-based K+ sensor. Pseudo-colored ratio images of HeLa cells expressing NES-GEPII 1.0 were treated with 15 µM gramicidin, a K+ selective channel forming peptide, in the absence or presence of 140 mM extracellular K+ as indicated. White numbers in the images represent the time in min:sec of image acquisition. Scale bar in the right lower image represents 10 µM.
Cells 08 00492 g003
Figure 4. Co-expression of genetically encoded metal ion indicators allows the correlation of intracellular Ca2+ with K+ dynamics on the single cell level. (a) Representative single cell response of an INS-1 832/13 cell, a pancreatic β-cell, expressing lc-LysM GEPII 1.0 (blue curve, both panels), a genetically encoded, FRET-based K+ indicator, and CAR-GECO1, a far-red fluorescent, single FP-based Ca2+ sensor (red curve, both panels). At the time point indicated in the panels, cells were depolarized by the application of 70 mM K+. The right panel demonstrates a zoom of the area indicated in the left panel. Single time-points measured during the depolarization are indicated (right panel, blue and red circles). (b) Analysis of the on- (left panel) and off-kinetics (right panel) of the intracellular K+ and Ca2+ transient upon cell depolarization as demonstrated in (a).
Figure 4. Co-expression of genetically encoded metal ion indicators allows the correlation of intracellular Ca2+ with K+ dynamics on the single cell level. (a) Representative single cell response of an INS-1 832/13 cell, a pancreatic β-cell, expressing lc-LysM GEPII 1.0 (blue curve, both panels), a genetically encoded, FRET-based K+ indicator, and CAR-GECO1, a far-red fluorescent, single FP-based Ca2+ sensor (red curve, both panels). At the time point indicated in the panels, cells were depolarized by the application of 70 mM K+. The right panel demonstrates a zoom of the area indicated in the left panel. Single time-points measured during the depolarization are indicated (right panel, blue and red circles). (b) Analysis of the on- (left panel) and off-kinetics (right panel) of the intracellular K+ and Ca2+ transient upon cell depolarization as demonstrated in (a).
Cells 08 00492 g004
Figure 5. Correlation of subcellular Ca2+ dynamics using genetically encoded probes and Fura-2. (a) Fluorescence wide-field microscopy images of HeLa cells loaded with Fura-2 (blue, left image) and expressing a FRET-based, red-shifted Ca2+ sensitive probe targeted to the lumen of the ER (D1ERCmR2, red, middle image). Right image shows the overlay of both fluorescence images. Scale bar represents 10 µM. (b) Fluorescence wide-field microscopy images of a HeLa cell loaded with Fura-2 (blue, left image) and expressing a FRET-based, red-shifted Ca2+ sensitive probe localizing in the mitochondrial matrix (4mtD1GO-Cam, red, middle image). Right image shows the overlay of both fluorescence images. Scale bar represents 10 µM. (c) Representative single cell response of a HeLa cell expressing D1ERCmR2 (red curve, both panels), additionally loaded with Fura-2 (blue curve, both panels) upon cell stimulation with Histamine and BHQ, in the absence (EGTA) or presence of Ca2+ as indicated in the panels. Right panel shows a zoom as indicated in the left panel, demonstrating the correlation of Ca2+ in the cytosol and endoplasmic reticulum. (d) Representative single cell response of a HeLa cell expressing mtD1GO-Cam (green curve, both panels), additionally loaded with Fura-2 (blue curve, both panels) upon cell stimulation with Histamine as indicated in the panels. Right panel shows a zoom as indicated in the left panel, demonstrating the correlation of Ca2+ in the cytosol and the mitochondrial matrix.
Figure 5. Correlation of subcellular Ca2+ dynamics using genetically encoded probes and Fura-2. (a) Fluorescence wide-field microscopy images of HeLa cells loaded with Fura-2 (blue, left image) and expressing a FRET-based, red-shifted Ca2+ sensitive probe targeted to the lumen of the ER (D1ERCmR2, red, middle image). Right image shows the overlay of both fluorescence images. Scale bar represents 10 µM. (b) Fluorescence wide-field microscopy images of a HeLa cell loaded with Fura-2 (blue, left image) and expressing a FRET-based, red-shifted Ca2+ sensitive probe localizing in the mitochondrial matrix (4mtD1GO-Cam, red, middle image). Right image shows the overlay of both fluorescence images. Scale bar represents 10 µM. (c) Representative single cell response of a HeLa cell expressing D1ERCmR2 (red curve, both panels), additionally loaded with Fura-2 (blue curve, both panels) upon cell stimulation with Histamine and BHQ, in the absence (EGTA) or presence of Ca2+ as indicated in the panels. Right panel shows a zoom as indicated in the left panel, demonstrating the correlation of Ca2+ in the cytosol and endoplasmic reticulum. (d) Representative single cell response of a HeLa cell expressing mtD1GO-Cam (green curve, both panels), additionally loaded with Fura-2 (blue curve, both panels) upon cell stimulation with Histamine as indicated in the panels. Right panel shows a zoom as indicated in the left panel, demonstrating the correlation of Ca2+ in the cytosol and the mitochondrial matrix.
Cells 08 00492 g005
Table 1. Selection of FRET-based GEPs suitable for subcellular K+ imaging.
Table 1. Selection of FRET-based GEPs suitable for subcellular K+ imaging.
GEPLocalizationKD In Vitro/In SituDynamic
Range
λExc
(nm)
λEm
(nm)
Ref.
GEPII 1.0Cytosol, Nucleus, Mitochondria, Subplasma membrane *420 µM/820 µM220%430475/525[30]
lc-LysM GEPII 1.030.47 mM/60.95 mM80%430475/525[30]
GEPII 2.7Cytosol3.24 mM/9.09 mM110%430475/525[30]
GEPII 2.104.39 mM/10.11 mM150%430475/525[30]
GEPII 2.158.59 mM/15.59 mM150%430475/525[30]
lc-LysM R-GEPII 1.0-/75.12 mM30%480510/560Unpublished
R-GEPII 1.0-/3.25 mM20%477510/560Unpublished
KIRIN11.66 mM/-130%430475/525[66]
KIRIN1-GR2.56 mM/-20%480510/560[66]
* indicates that the probes were fused to specific targeting sequences to obtain the given localizations, respectively.
Table 2. Selection of FRET-based GEPs suitable for subcellular Mg2+ imaging.
Table 2. Selection of FRET-based GEPs suitable for subcellular Mg2+ imaging.
GEPLocalizationKDDynamic
Range
λExc
(nm)
λEm
(nm)
Ref.
MagFRET-1Cytosol150 µM49%430475/525[34]
MagFRET-2350 µM33%430475/525[34]
MagFRET-39.2 mM58%430475/525[34]
MagFRET-48.5 mM62%430475/525[34]
MagFRET-57.4 mM74%430475/525[34]
MagFRET-615 mM50%430475/525[34]
MagFRET-7780 µM38%430475/525[34]
MagFRET-8890 µM56%430475/525[34]
MARIO7.2 mM153%430475/525[88]
MagFRET-1 NLSNucleus150 µM49%430475/525[34]
NLS-MARIO7.2 mM153%430475/525[88]
Table 3. Selection of FRET-based GEPs suitable for subcellular Ca2+ imaging.
Table 3. Selection of FRET-based GEPs suitable for subcellular Ca2+ imaging.
GEPLocalizationKDDynamic
Range
λExc
(nm)
λEm
(nm)
Ref.
YC2.1Cytosol100 nM and 4.3 µM2430475/525[27,129]
YC3.11.5 µM2430475/525[27,129]
YC3.31.5 µM~ 1.1430475/525[27,115]
YC6.1110 nM2430475/525[116]
YC3.60250 nM560%430475/525[117]
D1800 nM and 60 µM-430475/525[118]
D2cpV30 nM and 3 µM5.3430475/525[119]
D3cpV600 nM5.1430475/525[119]
D4cpV64 µM3.8430475/525[119]
TN-humcTnC470 nM100%430475/525[123]
TN-L151.2 µM100%430475/525[123]
TN-XL2.5 µM400%430475/525[124]
LynD3cpVSubplasma-membrane600 nM5.1430475/525[119]
Cav2.2-TN-XL2.5 µM-430475/525[122]
TN-L15D107ARas29 µM100%430475/525[123]
H2BD1cpVNucleus800 nM and 60 µM-430475/525[120]
4mtD3cpVMitochondria600 nM5.1430475/525[119]
4mtD1GO-CamMitochondria1.53 µM-477510/560[48]
N33D1cpVOuter mitochondrial-membrane800 nM and 60 µM-430475/525[120]
Split YC7.3erEndoplasmic Reticulum130 µM-430475/525[121]
D1ER220 µM-430475/525[118]
D1ERCmR2200 µM-480510/560[32]
apoK1-er124 µM-430475/525[126]
YC4.3ER800 nM and 700 µM-430475/525[118]
Table 4. Selection of FRET-based GEPs suitable for subcellular Cu+ imaging.
Table 4. Selection of FRET-based GEPs suitable for subcellular Cu+ imaging.
GEPLocalizationKDDynamic
Range
λExc
(nm)
λEm
(nm)
Ref.
Amt1-FRETCytosol2.5 aM-430475/525[36]
Ace1-FRET4.7 aM-430475/525[142]
Mac1-FRET97 zM-430475/525[142]
eCALWY-C2M/C3M--430475/525[153]
PMtb-FRET-3.31 zM-430475/525[146]
Table 5. Selection of FRET-based GEPs suitable for subcellular Zn2+ imaging.
Table 5. Selection of FRET-based GEPs suitable for subcellular Zn2+ imaging.
GEPLocalizationKDDynamic
Range
λExc
(nm)
λEm
(nm)
Ref.
CA + WYCytosol350 pM-430475/525[35]
Cys2His2Cytosol, Mitochondria *1.7 µM2.2430475/525[163]
His4Cytosol, Mitochondria, Plasma membrane *160 µM4430475/525[163]
eCALWY-1Cytosol, Insulin-storing granules *2 pM2430475/525[159]
eCALWY-2Cytosol9 pM2430475/525[159]
eCALWY-3Cytosol45 pM1.7430475/525[159]
eCALWY-4Cytosol630 pM2430475/525[159]
eCALWY-5Cytosol1.8 nM1.8430475/525[159]
eCALWY-6Cytosol, Insulin-storing granules *2.9 nM1.8430475/525[159]
ZinCh-6-260 nM~ 3.5430475/525[168]
ZinCh-9-500 nM and 88 µM4430475/525[168]
ZapCY1ER, Golgi Apparatus *2.5 pM4.15430475/525[170]
ZapSM2Cytosol, Nucleus *-1.1400510/560[171]
ZapSR2Cytosol, Nucleus *-1.2400510/580[171]
ZapOC2Cytosol, Nucleus *-1.1550565/610[171]
ZapOK2Cytosol, Nucleus *-1.1550565/635[171]
ZapCmR1Cytosol, Nucleus *-1.15480510/560[171]
ZapCmR1.1Cytosol, Nucleus *-1.5480510/560[171]
ZapCmR2Cytosol, Nucleus *-1.4480510/560[171]
eZinCh-2Cytosol, Mitochondria,
ER, Vesicles *
1 nM300%430475/525[169]
* indicates that the probes were fused to specific targeting sequences to obtain the given localizations, respectively.

Share and Cite

MDPI and ACS Style

Bischof, H.; Burgstaller, S.; Waldeck-Weiermair, M.; Rauter, T.; Schinagl, M.; Ramadani-Muja, J.; Graier, W.F.; Malli, R. Live-Cell Imaging of Physiologically Relevant Metal Ions Using Genetically Encoded FRET-Based Probes. Cells 2019, 8, 492. https://0-doi-org.brum.beds.ac.uk/10.3390/cells8050492

AMA Style

Bischof H, Burgstaller S, Waldeck-Weiermair M, Rauter T, Schinagl M, Ramadani-Muja J, Graier WF, Malli R. Live-Cell Imaging of Physiologically Relevant Metal Ions Using Genetically Encoded FRET-Based Probes. Cells. 2019; 8(5):492. https://0-doi-org.brum.beds.ac.uk/10.3390/cells8050492

Chicago/Turabian Style

Bischof, Helmut, Sandra Burgstaller, Markus Waldeck-Weiermair, Thomas Rauter, Maximilian Schinagl, Jeta Ramadani-Muja, Wolfgang F. Graier, and Roland Malli. 2019. "Live-Cell Imaging of Physiologically Relevant Metal Ions Using Genetically Encoded FRET-Based Probes" Cells 8, no. 5: 492. https://0-doi-org.brum.beds.ac.uk/10.3390/cells8050492

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop