Skip to main content

Regulation of mitochondrial functions by protein phosphorylation and dephosphorylation

Abstract

The mitochondria are double membrane-bound organelles found in most eukaryotic cells. They generate most of the cell’s energy supply of adenosine triphosphate (ATP). Protein phosphorylation and dephosphorylation are critical mechanisms in the regulation of cell signaling networks and are essential for almost all the cellular functions. For many decades, mitochondria were considered autonomous organelles merely functioning to generate energy for cells to survive and proliferate, and were thought to be independent of the cellular signaling networks. Consequently, phosphorylation and dephosphorylation processes of mitochondrial kinases and phosphatases were largely neglected. However, evidence accumulated in recent years on mitochondria-localized kinases/phosphatases has changed this longstanding view. Mitochondria are increasingly recognized as a hub for cell signaling, and many kinases and phosphatases have been reported to localize in mitochondria and play important functions. However, the strength of the evidence on mitochondrial localization and the activities of the reported kinases and phosphatases vary greatly, and the detailed mechanisms on how these kinases/phosphatases translocate to mitochondria, their subsequent function, and the physiological and pathological implications of their localization are still poorly understood. Here, we provide an updated perspective on the recent advancement in this area, with an emphasis on the implications of mitochondrial kinases/phosphatases in cancer and several other diseases.

Background

The mitochondria are double membrane-bound organelles found in most eukaryotic cells. They generate most of the cell’s energy supply of adenosine triphosphate (ATP). Protein phosphorylation/dephosphorylation is a crucial regulatory system of signal transduction which controls many aspects of cellular functions. It can greatly impact the properties of a protein’s enzymatic activity, structure, subcellular localization and stability. Phosphorylation, accomplished via a kinase action, and its reverse-de-phosphorylation via a phosphatase are essential switching-on and -off mechanisms in cell signaling. Many kinases and phosphatases have been well-documented to play roles in the cell primarily outside mitochondria. Although the function and mechanisms of translocation to mitochondria are not well understood, the evolution of imaging and molecular techniques has allowed the discovery that an increasing number of kinases and phophatases localize to mitochondria. The importance of phosphorylation/dephosphorylation in the governing of mitochondrial processes is supported by recent evidence that mitochondria contain many protein kinases, phosphatases, and associated proteins.

The aim of this review is to provide an updated perspective on the kinases and phosphatases localized to mitochondria. We describe here kinases and phosphatases that associate with mitochondria, their function within mitochondria, identified thus far, and the possible physiological or pathological implications of their mitochondrial localization.

Kinases localized in mitochondria

Tyrosine kinases

Abl

Abl is a non-receptor tyrosine kinase and is ubiquitously expressed in several cell types and tissues. Abl is localized at distinct sub-cellular sites including the nucleus, cytoplasm and endoplasmic reticulum (ER) and has various binding partners including signaling adaptors, proliferation regulators, transcription factors, kinases, phosphatases and cytoskeletal proteins [1]. While Abl is primarily localized to the nucleus, ER, and cytoplasm, it has been found that about 4 % of the total Abl resides in the mitochondria of murine embryo fibroblasts [2]. It has been revealed that H2O2 treatment of various cell types causes translocation of Abl to mitochondria and then induces cell death. Apart from H2O2, stimuli of ER stress, calcium ionophore A23187 and brefeldin A, also promote the translocation of Abl to mitochondria [3]. Although Abl does not contain a typical mitochondrial localization signal of its own, a study showed that protein kinase Cδ binds to Abl in the ER. This PKCδ-Abl complex translocates from ER to mitochondria and then triggers apoptosis [4].

Abl contributes to human diseases such as Alzheimer’s disease (AD) and chronic leukemia. Exposure of cultured neurons to various forms of multimeric amyloid-β peptide led to an increase in Abl tyrosine kinase activity and subsequent tyrosine phosphorylation of tau at Y394 [5–7]. Additionally, it was found that abnormal tau phosphorylation and subsequent cell death could be prevented by the Abl family kinase inhibitor [6, 7].

In chronic myeloid leukemia (CML) and acute promyelocytic leukemia (AML), Abl is activated upon translocation within breakpoint cluster region (Bcr) gene [8, 12]. Bcr/Abl fusion protein is a constitutively active tyrosine kinase which allows cells to proliferate without cytokine regulation, leading to a clonal myeloproliferative disorder. Its inhibitor, imatinib mesylate, can trigger apoptosis of CML cells [9]. It will be intriguing to explore if Abl or Bcr/Abl-associated apoptosis in AD and chronic leukemia is directed by mitochondrial localized Abl activity.

Src family kinases

There are multiple lines of evidence support the mitochondrial localization and function of Src family kinases. Although Src family kinases are permanent residents of cytoplasm, a previous study has suggested Src family including Src, Fyn, Lyn, Fgr and a negative regulator of Src family kinases, CSK, can be localized in mitochondria [10]. The phosphotyrosine signal in mitochondria was dramatically decreased by pretreatment with a Src inhibitor, PP2. This result suggests that Src family kinases may be associated with mitochondrial protein phosphorylation. Other groups performed subfractionation of mitochondria using proteinase K and/or Triton X-100, further supported the localization of Src family kinases within mitochondria [11].

Since Src family kinases do not contain typical mitochondrial localization signals, they seem to be dependent on unidentified adaptor proteins for translocation into mitochondria. Several studies revealed that two anchoring proteins of protein kinase A (PKA), A Kinase Anchor Proteins 121 (AKAP121) and Dok-4, associate with Src. It is well known that AKAP121 acts as a multifunctional protein binding to protein tyrosine phosphatase D1 (PTPD1). PTPD1 anchors Src to the outer membrane of mitochondria upon activating the protein [12, 13]. Interestingly, a study showed that AKAP121 is found in the mitochondrial inner membrane with Src and Lyn [14]. Dok-4 is an adapter protein and appears to be responsible for the mitochondrial importation of Src in bovine endothelial cells. While deficiency of Dok-4 induces Src localization to extra-mitochondria, overexpression of Dok-4 causes Src localization to mitochondria [15].

Mitochondrial electron transport chain (ETC) is the final component of aerobic respiration, which contains a series of electron transporters embedded in the inner mitochondrial membrane. It includes complex I to IV and creates a chemical gradient that allows for the production of ATP. Several ETC complexes have been identified as substrates for Src. The first protein shown to be activated by mitochondrial Src was cytochrome c oxidase (complex IV in ETC) [11]. Src phosphorylates a subunit of cytochrome c oxidase, and leads to the enzymatic activation of complex IV in osteoclasts. Recently, it has been found that Src also affects other complexes within the ETC. One study showed that an increase in Src activity, as a response to changes in ATP levels in mitochondria of the rat brain, leads to an increase in the activities of complexes I, III, and IV, along with decreases in activity of complex V [16]. Two groups reported that inhibition of Src activity led to reduction in mitochondrial respiration mediated by a specific decrease in actions of complex I’s NADH dehydrogenase-ubiquinone oxidoreductase system. NADH ubiquinone flavoprotein 2 at Tyr 193 of complex I and succinate dehydrogenase A at Tyr 215 of complex II, are the targets of Src phosphorylation [17, 18].

c-Src, normal proto-oncogene Src, possesses anti-apoptotic properties and shows increased protein levels and activity in a variety of human cancers [19]. Further, c-Src has been found to be a crucial player in multiple signaling pathways regulating proliferation, survival, metastasis, and angiogenesis [19]. One of these pathways is that of the mitochondrial, hypoxia-mediated ROS-induced activation of c-Src, HIF-1α, and NF-κB which contributes to cell survival in various cancers including hepatomas, colon carcinomas, and neuroblastomas [20]. Additionally, c-Src has an important role in serum deprived pancreatic cancer cell survival [21]. Also, c-Src leads to metastasis of diverse cancers including breast cancer [22].

Mitochondrial Fgr kinase regulates complex II activity by tyrosine phosphorylation and plays an important role in regulating NADH/FADH2 ratio. H2O2 promotes Fgr localization to mitochondria, but it does not seem that Fgr has a consensus signal sequence for mitochondrial localization [23].

Hibbs et al. found that mice homozygous for Lyn locus aberrations show abnormalities connected to B lymphocyte lineage and mast cell function [24]. Lyn-/- mice did not mediate an allergic response to IgE cross-linking and exhibited severe glomerulonephritis caused by the kidney deposition of IgG immune complexes [24, 25], indicating that Lyn is associated with autoimmune disease. Lyn was also discovered to be critical for maintenance of the leukemic phenotypes of many different hematopoietic cancers including AML, CML and B cell lymphocytic leukemia [26–28]. Lyn was also expressed in some solid tumors. Therefore, it could serve as a potential therapeutic target for prostate cancer, glioblastomas, and aggressive subtypes of breast cancer [29–31]. Recent study revealed that Lyn-mediated mitochondrial tyrosine phosphorylation is required for hepatocyte survival under partial hepatectomy, pro-apoptotic conditions [32]. However, the role of Lyn in mitochondria and human diseases needs to be further elucidated.

Receptor tyrosine kinases

EGFR

A recent study demonstrated that epidermal growth factor receptor (EGFR) is localized to mitochondria where it interacts with cytochrome c oxidase subunit II (CoxII) in mouse fibroblasts over-expressing EGFR and Src [33]. EGF stimulation promotes translocation of EGFR into mitochondria. The translocation depends upon the phosphorylation of EGFR on Y845 by Src. The translocation is not observed upon overexpression of a catalytically inactive Src mutant or upon replacement of Y845 by phenylalanine (Y845-EGFR). A similar study showed that CoxII was phosphorylated by Src and EGFR, and the EGFR’s mitochondrial localization was increased by EGF stimulation. Mitochondrial EGFR decreased the activity of Cox and reduced cellular ATP, indicating that mitochondrial EGFR regulates the function of mitochondria [34]. Moreover, new evidence showed that EGFR and EGFRVIII, a constitutively active variant of EGFR were localized in mitochondria [35, 36]. EGFR co-localized with FAK and Src in the mitochondria in glioma and xenografts. This suggests a role of mitochondrial EGFR in cancer progression [37].

There is a debate as to whether the translocation of EGFR depends upon its endocytosis. One group concluded that the EGFR’s localization to mitochondria depends on its endocytosis [38]. The report revealed that EGFR’s mitochondrial translocation can be dramatically decreased in the presence of 3-methyladenine (3-MA), an autophagy inhibitor. Also, the knockdown of Beclin 1, an autophagy related gene, markedly decreased the mitochondrial translocation of EGFR. However, another group suggested that mitochondria-localized EGFR is independent of its internalization [39]. Therefore, the detailed mechanisms of EGFR mitochondrial localization still remain unclear.

ErbB2

ErbB2 is a human version of EGFR2. Recent study revealed that ErbB2 localizes to mitochondria of breast cancer cells and tumor samples of patients [40]. The study also found that localization of ErbB2 into mitochondria is mediated via association with mtHSP70 and reduces mitochondrial respiratory functions, including oxygen consumption, while it enhanced cellular glycolysis. It also showed that mitochondrial ErbB2 decreased membrane potential as well as the cellular ATP levels. However, detailed molecular mechanisms or direct substrates of mitochondrial EGFR and ErbB2 remain unclear. It is well known that EGFR and ErbB2 have been associated with the development of numerous human cancers. Tumors with changes in EGFRs tend to be more aggressive, and are considered as indicators of a poor clinical outcome, therefore they are intensely studied as therapeutic targets [41].

Serine/threonine kinases

Akt

It is well known that activated protein kinase B (Akt) is localized to diverse subcellular compartments: these include the Golgi, endoplasmic reticulum, and the nucleus. Also, it was shown that Akt could be localized in mitochondria when stimulated by 17β-estradiol and insulin in endothelial cells [42]. In human neuroblastoma, it was reported that the activation of PI3 K/Akt signaling via insulin or insulin-like growth factor-1, is greatly amplified by mitochondrial Akt locolization [43]. Recent studies also showed that treatment with leukemia inhibitory factor (LIF) increased the amount of total and phospho-Ser473-Akt in mitochondria [44], and insulin treatment could translocate Akt to mitochondria in cardiac muscle cells [45].

Several studies reported that mitochondrial Akt causes phospho-inactivation of the pro-apoptotic protein BAD, and recruitment of Raf-1 to the mitochondria promoted cell survival [46, 47]. Recent studies also discovered new targets for mitochondrial Akt such as mitochondrial electron complex V and hexokinase-II [44, 45]. Akt isoform 1 modulates mitochondrial complex V activity, enhances the production of ATP, and increases phosphocreatine in cardiomyocytes, indicating that Akt is associated with ATP metabolism [45]. Akt also has a direct effect in mitochondria, which is mediated by phosphorylation of hexokinase-II, resulting in protection of the mitochondria from oxidant or Ca2+-induced mitochondrial permeability transition pore (MPTP) opening [44]. Yang et al. found that translocation of phospho-Akt to mitochondria was enhanced in the streptozotocin-induced diabetic mice and insulin stimulates translocation of Akt to mitochondria [48]. Barksdale et al. revealed that HSP90 is responsible for Akt accumulation in mitochondria in unstimulated cells [49]. Taken together, mitochondrial Akt may play important roles in energy metabolism associated with diabetes.

JNK

It was well established that the c-Jun N-terminal kinase (JNK) localizes in cytoplasm and the nucleus. Many studies show that JNK may be activated in or translocated to the mitochondria. It was first revealed that JNK is localized in mitochondria in the phorbol ester response of myeloid leukemia cells [50]. Recent studies suggested an important role of JNK in mitochondria-related functions. These studies include murine heart mitochondria, hydrogen peroxide-treated rat brain or primary cortical cultures, acetaminophen-induced liver damage, multiple myeloma cells treated with anti-cancer drugs, and neonatal ischemia [51–57].

As mentioned before, JNK regulates apoptosis in general, indicating that targets of JNK can be apoptosis related proteins. JNK mediates phosphorylation and oligomerization of proapoptotic BAD, initiating apoptosis [58]. Conversely, there are studies reporting that JNK reduces apoptosis. For example, JNK is necessary for IL-3-mediated cell survival via phosphorylation and inactivation of BAD [59]. Also, activated JNK may co-localize with, and phosphorylate, Bcl-2 in mitochondrial membranes of hematopoietic cells [60]. This effect is present during interleukin-3-mediated stress, resulting in enhanced anti-apoptotic functions of Bcl-2. These contradictory results indicate that the localization and function of JNK in mitochondria are not well understood and warrant further investigation.

ERK1/2

Although extracellular signal-regulated kinase 1 and 2 (ERK1/2) normally translocates between cytosol and the nucleus to influence trophic and pro-survival functions, recent studies found that ERK1/2 can localize to mitochondria of mouse heart [51], renal epithelial cells [61], mouse hippocampus [53] and human alveolar macrophages [62]. Mitochondrial ERK1/2 appears to play a crucial part in mitochondrial function [61, 62] including mitochondrial dysfunction, mitophagy, and apoptosis [63–65]. Recent investigation found that mitochondrial ERK1/2 phosphorylates steroidogenic acute regulatory protein (StAR), and this phoshporylation by ERK is required for the maintenance of this protein in mitochondria. Mitochondrial StAR together with mitochondrial active ERK and PKA are necessary for maximal steroid production [66].

p38 MAPK

Several studies revealed that the p38 mitogen-activated protein kinase (p38 MAPK) could reside in mitochondria. One group, using immunoblotting and immunofluorescence, showed that p38 MAPK is localized in cardiac mitochondria [51]. p38 MAPK was activated in the mitochondrial fraction by ischemia [67] and increased in mitochondria under H2O2 treatment [68]. Although it is well known that p38 MAPK performs a crucial function in mitochondria-mediated apoptosis, the precise function of mitochondrial p38 MAPK remains to be elucidated.

GSK3β

It has been established that glycogen synthase kinase (GSK) 3β can be localized in cytosol and nucleus. Some substrates of GSK3β (e.g., tau), are cytosolic, whereas others (many transcription factors) are nuclear. However, in 1995, it was revealed that GSK3β is present in mitochondria in rat cerebellum [69]. Since then several studies showed that GSK3β regulates the function of mitochondrial proteins such as adenine nucleotide translocator (ANT) and cyclophilin D [70–72]. The mechanism of GSK3β’s translocation to the mitochondria is unclear, but TOM20 may be integral, as in the example of connexin-43 [73]. In mitochondria, GSK3β phosphorylates and suppresses a critical mitochondrial enzyme, pyruvate dehydrogenase activity [74]. Also, it has been demonstrated that GSK3β has a role in determining the threshold for mitochondrial permeability transition pore (MPTP) opening [75], indicating that GSK3β regulates mitochondria-mediated apoptosis.

Accumulated evidence shows that GSK3β is highly associated with neurodegenerative diseases. GSK3 is a key kinase contributing to aberrant phosphorylation of the microtubule-binding protein tau in a process thought to cause neurofibrillary tangles in Alzheimer’s disease [76]. In 1994, Mulot et al. found, via SDS-PAGE, that paired helical filament-tau from the AD brain comprises four species which can be mimicked by GSK3β mediated human brain tau phosphorylation in vitro [77]. Recent study also revealed that, in primary cortical neurons, GSK3β is involved in Amyloid-β phosphorylation [78]. GSK3β polymorphism was found to be associated with Parkinson’s disease (PD) that alters GSK3β transcription and splicing [79].

There is much interest in GSK3β as a potential therapeutic target in many types of cancer. However, the application is complicated by findings that GSK3β behaves as a tumor suppressor, but may promote cell proliferation in different types of cancer. There is evidence that GSK3β functions as a tumor suppressor in skin and breast cancer [80, 81] and that repression of RNA polymerase 1 transcription by GSK3β contributes to this tumor suppressor action [82]. However, there are contradictory findings that GSK3β mediates tumor promotion and/or GSK3β shows anti-proliferative effects in certain types of tumors including colon and pancreatic cancer [83, 84]. Whether these appeared contradictory functions of GSK3β can be explained by the different localizations of this protein? It is still not clear if mitochondrial GSK3β is associated with neurodegerative diseases or cancer.

PKA

It has been acknowledged for many years that Protein kinase A (PKA) is crucial to mammalian mitochondrial physiology. Since the 1970s, it has been shown that PKA associates with, and in, mitochondria [85–88]. While catalytic subunits of PKA are predominantly found in the mitochondrial outer membrane [89], there is much evidence that PKA localizes to the mitochondrial inner membrane and matrix as well [87, 90–93]. In general, PKA is concentrated in cellular membranes and organelles through interactions with A kinase anchor proteins (AKAPs) [94, 95]. One group showed that de-localization of PKA from mitochondria was prompted by dominant negative AKAP121 mutation [96]. Indeed, expression of AKAP121 targets PKA to the cytoplasmic surface of mitochondria and then refines cAMP-PKA signaling to mitochondria [97]. Recent investigation revealed that AKAP1 recruits PKA and other signaling proteins to the outer mitochondrial membrane and thereby integrates several second messenger cascades by inhibitory phosphorylation of dynamin-related protein 1 (Drp1) and maintenance of mitochondrial integrity [98].

Drp1 is a mechanoenzyme using GTP hydrolysis to fuel the division of mitochondria. PKA mediated phosphorylation of Ser637 on Drp1 blocks Drpl’s translocation to the mitochondrial surface and influences on fission [99]. Other findings indicate that outer mitochondrial PKA, and phosphatase PP2A, regulate neuronal development by inhibiting and promoting mitochondrial division [100].

The cAMP response element-binding protein (CREB) is a transcription factor which regulates the transcription of cAMP response element-regulated genes and is found to bind to the mitochondrial DNA [101]. It has been shown that CREB is involved in regulation of mitochondrial gene expression and neuron longevity. Interestingly, it has been suggested that CREB may be phosphorylated by PKA within mitochondria [102].

Many heat shock proteins (Hsps) are known to play essential protective roles in the cardiovascular system. Hsp20 is expressed at high levels in cardiac, skeletal, and vascular smooth muscle. Hsp20 is regulated by the β-adrenergic/cAMP/PKA signaling pathway, which is known to be chronically activated in heart failure [103]. A recent study also revealed that PKA’s phosphorylating Ser16 of Hsp20 is vital to the small Hsp’s cardioprotective action [104]. For therapeutic development, it will be critical to distinguish if PKA mediated cardiprotective effects are due to mitochondrial PKA activity.

PKC

Protein kinase C (PKC) is a group of phospholipid-dependent serine/threonine kinases regulating numerous cellular functions through key signaling molecules. The activated PKCs translocate to multiple subcellular sites. Interestingly, analysis of the subcellular distribution of PKCε in mouse heart with constitutively active PKCε revealed that activated PKCε is associated with a variety of mitochondrial proteins [51]. It has been suggested that mitochondrial translocation of PKCε is associated with cardioprotection [105], and several studies showed that PKCε substrates reside within cardiac mitochondria [106–108]. COX IV has been identified as a substrate of PKCε in mitochondria [106]. Phosphorylation and activity of COX IV were increased by PKC activator, phorbol 12-myristate 13-acetate (PMA), but were blocked by the selective PKCε inhibitor, εV1-2. Another known mitochondrial target for PKCε is the MPTP [107]. MPTP remains tightly closed under normal conditions, but it opens under cell death conditions. PKCε interacts with presumed components of the MPTP in heart mitochondria, leading to reduced MPTP opening. Another target for PKCε is a mitochondrial ATP-sensitive K+ channel (mitoKATP) [108]. It was recently found that PKC mediates interactions between conexin43 and mitoKATP subunit in mitochondria of cardiomyocytes [109]. These channels are normally closed, but are opened during periods of metabolic stress as ATP levels decline, leading to channel opening-mediated protective effects. This evidence reveals that mitochondrial PKCε has a pro-survival role in a stress condition.

Hamasaki et al. found that expression of PKCε decreases by 40 % in cardiac ventricles of hyperthyroid rats [110]. The expression of PKCε was reduced in cytosol and membrane fractions but was not reduced in extracts of hypertrophied cardiac ventricles generated by aortocaval shunt or aortic banding. Other groups also revealed that hypertrophic stimuli activated PKCε in rat-cultured cardiac myocytes, and in vivo, and that PKCε overexpression in mice leads to cardiac hypertrophy related to concentric remodeling and the preservation of cardiac contractility [111, 112].

Recent study revealed that PKCɛ increases the activity of endothelin converting enzyme, which degrades amyloid-β, and decreases amyloid plaque in transgenic mice [113], indicating that greater neuronal PKCɛ activation may promote amyloid-β clearance, reducing Alzheimer’s neuropathology. However, it has not been examined whether mitochondrial PKCɛ is needed for AD progression.

PKCδ also translocates to mitochondria. Diverse studies revealed that treatment in a variety of neoplastic cells with hydrogen peroxide, phorbol esters, or anticancer agents such as cisplatin and etoposide causes PKCδ translocation into the mitochondria [114–116]. A study also found that mitochondrial PKCδ triggers the release of cytochrome c and resultant cell death. Thus, overexpressed PKCδ promotes apoptosis in neoplastic and normal keratinocytes by targeting mitochondria and disturbing their membrane potential [117]. However, direct targets for mitochondrial PKCδ still remain unclear.

PINK1

The PTEN-induced kinase 1 (PINK1) is a serine/threonine kinase. It is well known that PINK1 is a factor in autosomal recessive familial PD [118]. A large amount of evidence shows that PINK1 is localized in mitochondria and regulates mitochondrial function. A study found a strong mitochondrial targeting signal domain at its N terminus and showed that N-myc-tagged PINK1 expressed in mammalian cells accumulated in mitochondria. The protein could be found on the inner and outer mitochondrial membrane [119, 120]. PINK1 regulates Parkin (a protein related to PD), which acts as an E3 ubiquitin ligase [121]. PINK1 recruits Parkin to depolarized mitochondria. While PINK1 is imported and rapidly degraded by presenilin-associated rhomboid-like serine protease (PARL) in mitochondria with intact membrane potential, PINK1 breakdown is impaired in mitochondria with reduced membrane potential [120]. Apart from Parkin, PINK1 phosphorylates the mitochondrial fusion protein MFN2, which then acts as a mitochondrial receptor for Parkin [122] and Miro, an atypical Rho GTPase that tethers mitochondria to the tubulin network [123]. Several studies showed that the PINK1/Parkin pathway is associated with fission and fusion events of mitochondria. Those studies revealed that the PINK1/Parkin in mitochondria regulates pro-fusion pathway given that the overexpression of PINK1 leads to interconnected and structurally elongated mitochondria, whereas knockdown of PINK1 leads to fragmentation of mitochondria [124, 125].

PINK1/Parkin pathway also regulates transport of mitochondria. Damaged mitochondria, prior to their clearance, are sequestered via the PINK1/Parkin pathway where mitochondrial movement is prevented through PINK1 mediated-Miro phosphorylation [123, 126]. PINK1 regulates mitochondrial content through mechanisms independent of mitophagy and PINK1-knockout mice showed a decreased Complex I activity [127, 128]. PINK1 mutant flies showed complex I deficiency and impaired ATP generation; however, supplementing these PINK1-deficient flies with complex 1 or ubiquinone corrects reductions in ATP generation allowing for recovery of flight muscle [129]. In both PINK1 and Parkin null flies, there lies a widespread defect with turnover of electron transport chain proteins [130]. Recent study demonstrated that PINK1 and Parkin mediate localized translation of respiratory chain component mRNAs along the mitochondrial outer membrane [131].

Dagda et al. studied the relationship between a PINK1 knock down and mitochondrial dysfunction in PD [132]. Stable knockdown of PINK1 produced mitochondrial fragmenting and autophagy in SH-SY5Y cells. Recent study found that loss of PINK1 accelerates neurodegenerative phenotypes induced by mitochondrial stress in transgenic animals where exists the conditional expression of mitochondrial unfolded ornithine transcarbamylase [133]. Mutations associated with PD were found throughout PINK1; however, the majority lay within the kinase domain, suggesting that loss of PINK1 kinase function is part of the pathogenesis [134]. Further, mutations within the carboxy terminus of PINK1 turn out to be critical to control an optimal kinase activity [135].

There is substantial evidence, based on epidemiological studies, for lower rates of cancer in patients with PD. The relationship between cancer incidence and PD may be related to the presence of genes common to both diseases [136]. Investigators found that PTEN knock-down suppresses PINK1 levels and shows cardioprotective effects, and concluded that PINK1 is a potential novel cardioprotective kinase [137]. However, the role of PINK1 in cancer development and progression is unclear. One study reported a link between Type II diabetes and PINK1 [138]. Investigators found that PINK1 was reduced in the skeletal muscle of Type II diabetic patients and in conditions of obesity or inactivity (or both) in patients and controls. One study found glucose transporter inhibition under loss of PINK1 [139]. Recently, Deas et al. reported that the loss of PINK1 function seems to disturb glucose-sensing, resulting in enhanced insulin release, which is uncoupled from glucose uptake in beta cells [140], suggesting that a deficiency of the Parkinson’s-associated PINK1 protein may directly alter insulin secretion at the pancreas.

Phosphatases translocate into mitochondria

Protein phosphatases

MKP1

MAP Kinase Phosphatases (MKPs) are dual specificity phosphatases which dephosphorylate both pSer/pThr and pTyr. MKP1 is localized in the nucleus through a LXXLL motif and appears to be confined to the dephosphorylation of nuclear MAPKs [141]. However, recent study found that MKP1 translocates to mitochondria after nerve growth factor treatment [142]. This result appears reasonable given that its substrates, and MAPKs, are also localized in mitochondria.

Wang et al. found that, in malignant samples, MKP1 is increased five fold over non-malignant samples [143]. This indicates that therapeutic suppression of MKP1 action may allow for expression of pro-apoptotic signaling of JNK in malignant cells. MKP1 is highly expressed in various human tumors, including non-small-cell lung cancer (NSCLC), bladder, ovarian, breast, osteosarcoma and prostate cancers [144–146]. They also found that MKP1 plays a critical part in the pathology of NSCLC, both in tumor growth and in response to treatment with cisplatin [146]. Recent study demonstrated that interference of MKP1 actions led to diminished invasion potential, tumor growth, and metastasis in mice [147].

A role for MKP1 in modulation of innate immune responses is strongly supported by several recent papers using models of endotoxic shock, anaphylaxis and arthritis. Mice deficient in MKP1 were hypersensitive to endotoxic shock; this was associated with prolonged activation of MAPKs, and enhanced production of TNF-α, IL-6, and IL-10 [148–150]. Moreover, it has been found that the absence of MKP1 markedly exacerbates disease development in mouse model of rheumatoid arthritis (RA). Overall, MKP1 controls dephosphorylation of MAPKs, but it is not clear whether this is the case for mitochondrial MAPKs and whether MKP1 is associated with human diseases due to mitochondrial MKP1 function.

Shp2

Src homology 2 domain-containing phosphatase 2 (Shp2) is ubiquitously expressed and contains two N-terminal SH2 domains along with a C-terminal protein tyrosine phosphatase domain. Shp2 was primarily located inside the mitochondria related to cristae and the intercristal space [16, 151]. Recent study also found that sepsis induced Shp2 mitochondrial localization and expression [152]. Mutations of Shp2 give rise to many distinct human diseases: Germ line Shp2 mutations result in Noonan Syndrome (NS), one of the most common autosomal dominant disorders, and LEOPARD Syndrome (LS), known, as its acronym reminds us, for its major manifestations: multiple Lentigines, Electrocardiographic aberrations, Ocular hypertelorism, Pulmonary stenosis, Abnormalities of genitalia, Retardation of growth, and sensorineural Deafness [153, 154]. The distinct pathogenic routes of NS and LS are driven by different Shp2 mutations: NS-associated Shp2 mutations typically alter residues at the interface of N-SH2 and PTP domains [155], leading to increased enzymatic activity and RAS/ERK activation. LS mutations impact residues of the PTP domain, resulting in dramatic reduction of catalytic activity along with reduced activation of RAS/ERK [156].

Somatic Shp2 mutations are found in ∼35 % of patients with sporadic juvenile myelomonocytic leukemia (JMML), a clonal pediatric myeloproliferative disorder (MPD) featuring the amplified expansion and tissue infiltration of myeloid cells, along with macrocytic anemia and persisting fetal hemoglobinemia [157, 158]. LS patients often undergo hypertrophic cardiomyopathy; in addition they may see an elevated potential for neuroblastoma and AML [159, 160]. Although shp2 expression and activity is closely associated with several diseases, connection between mitochondrial shp2 and disease progression has not been well established.

PTPs

Several protein tyrosine phosphases (PTPs) such as PTPD1, PTP1B, and PTPMT1 has been reported as mitochondrial localized PTPs. PTPD1 is localized to outer membrane of mitochondria by binding to AKAP121 (or spliced isoform AKAP84) and increase a magnitute of EGF stimulated signaling [13]. In addition, PTPD1 binds to Src and activate Src, enhancing EGF-dependent mitogen signaling, enhancing oxidative phosphorylation and mitochondrial membrane potential [12]. PTP1B also localizes to mitochondria and enhances Src mediated mitochondrial oxidative phosphorylation [161]. Dual phosphatase PTPMT1 (PTP localized to the mitochondrion 1) has N-terminal mitochondrial localization signal sequence and is found in the matrix face. Knock-down of PTPMT1 alters mitochondrial phosphoprotein profile and markedly enhances ATP production [162]. This study clearly demonstrated the importance of protein phosphorylation/dephosphorylation switch in regulating mitochondrial function.

Lipid phosphatase

PTEN

Phosphatase and tensin homolog deleted on chromosome 10 (PTEN) is a well-known lipid phosphatase. Early studies proposed that PTEN localized exclusively to the cytoplasm and was able to transiently associate with the plasma membrane depending on the local PIP2 and PIP3 concentrations [163]. However, PTEN has recently been shown to localize to specialized subcellular compartments, such as the nucleus and the nucleolus, the mitochondria and the endoplasmic reticulum. Zhu et al. observed that a gradual buildup of PTEN in mitochondria occurred after induction of apoptosis, which was accompanied by the translocation of Bax to mitochondria [164]. Further, Zu et al. revealed that ischemia/reperfusion (I/R) induces mitochondrial localization of PTEN in the myocardium. Moreover, ischemic preconditioning attenuates mitochondrial localization of PTEN post-I/R, possibly blocking the translocation of Bax to the mitochondria, and leading to improved cell viability [165]. Liang et al. characterized an N-terminally extended form of PTEN (named PTENα) that localizes to the cytoplasm and mitochondria, and induces cytochrome c oxidase activity and ATP generation in mitochondria [166]. This evidence showed that PTEN could exist in mitochondria, playing a crucial role in cellular functions such as apoptosis. However, whether this function of PTEN depends on its lipid or protein phosphatase activities and what are the substrates in mitochondria are still not clear.

Molecular machinery for protein import into mitochondria

The majority of mitochondrial proteins are encoded in the nucleus, synthesized in the cytosol, and then delivered into their proper organelle. These mitochondrial proteins are imported into one of four mitochondrial compartments: outer membrane, intermembrane space, inner membrane, and matrix (Fig. 1). Each compartment contains translocases, which interact with precursor proteins to regulate their transport. The major molecular machinery which translocates proteins across the mitochondrial outer membrane is TOM (translocase of the outer membrane of mitochondria) complex. TOM comprises numerous integral membrane protein components: receptor subunits including Tom70 and Tom20, and core translocase subunits including Tom5, Tom6, Tom7, Tom22, and Tom40 [167, 168]. Upon transiting the channel of the TOM complex, substrate proteins may interact with one of three distinct machineries: translocase of inner mitochondrial membrane (TIM) complex 23, Tim9-Tim10 chaperone complex, and mitochondrial intermembrane space assembly machinery (MIA). The classic mitochondrial protein import pathway involves N-terminal presequences on the precursor proteins. These proteins, following passage through the TOM complex, are directed to TIM23 complex and TIM23 bound presequence translocase-associated motor (PAM) which completes pre-protein translocation into the matrix. Following this, mitochondrial processing peptidases remove the presequences, and the proteins are folded to their native conformations. However, many mitochondrial precursor proteins are translated without cleavable presequences. These proteins also can be imported into mitochondria. The precursors of outer membrane beta-barrel proteins are transported, via the Tim9-Tim10 chaperone complex, to the sorting and assembly machinery (SAM) complex of the outer membrane [169]. Spanning proteins of the inner membrane also interact with the Tim9-Tim10 chaperone complex, after which they are inserted into the inner membrane by the TIM22 complex [170]. Several proteins of the intermembrane space contain cysteine motifs and are imported and oxidized by MIA [171].

Fig. 1
figure 1

Mitochondrial import pathways. First, nuclear-encoded mitochondrial proteins are imported by the TOM complex. Then, these proteins head to different pathways. Proteins with presequence are transported by the TIM23 complex and the PAM into the matrix. The intermembrane space localized proteins are imported via the MIA. β-barrel precursors of the outer membrane are transferred by the Tim9–Tim10 complex from TOM to SAM. Precursors of inner membrane carriers are inserted by the TIM22 complex into the inner membrane

As far, the detailed information of the mechanism of translocation of the kinases and phosphatases to mitochondria are elusive. The majority of the reported mitochondrial localized kinases/phosphatases lack typical mitochondrial localization signal. This raises three possibilities. First, it does not need a typical mitochondrial localization signal to be localized in mitochondria; second, there are hidden undiscovered mitochondrial localization signal; third, it is passively transported to mitochondria by a cargo carrier. Since it has been shown that mitochondrial kinases/phosphatases play important roles in regulating various cellular functions, more investigations on the mechanism of the importation are definitely needed.

Concluding remarks and future directions

This review summarizes recent findings concerning mitochondrial localization of kinases and phosphatases, recaps our knowledge on their function to cellular domains beyond those commonly cited in textbooks (Table 1). Over the past many years, the role of mitochondria has been underestimated and understudied. Evidence revealed by recent advanced laboratory techniques such as tandem mass spectrometry, confocal imaging, and electron microscopy, reevaluate mitochondria as a crucial platform for cellular signaling beyond just energy production in the cell. Mitochondria have diverse reported phosphoproteins, kinases and phosphatases, as we have described here. Mounting evidence support that phosphorylation or dephosphorylation of mitochondrial proteins influences mitochondrial function, including metabolism of sugar, amino acids, and lipids; oxidative phosphorylation, antioxidant protein expression, mitochondrial fission and fusion, and decision for survival/death. Emerging evidence has shown that mitochondrial kinases/phosphatases regulate a variety of key regulatory processes in diverse mammalian tissue [172–175]. Intriguingly, a recent phosphoproteome analysis of functional mitochondria showed there are 155 phosphorylation sites in 77 mitochondrial proteins including inner membrane ETS and enzymes in resting human muscle [176]. Although the detailed functions of these kinases/phosphatases in mitochondria are still unclear, there is evidence that the translocated mitochondrial kinases/phosphatases may be seen in mitochondria in several mitochondrial compartments: Those in the matrix likely regulate enzymes of the TCA cycle, amino acid metabolism, free-radical balance, ETC activity, and ATP synthesis (Fig. 2). Those in the outer membrane regulate transport, cell death-related proteins, and mitochondrial fission/fusion. Those at the inner membrane regulate free-radical balance, nucleotide transport, ETC activity and complex assembly.

Table 1 Mitochondrial kinases/phosphatases
Fig. 2
figure 2

Schematic depiction of the functions of mitochondrial kinases/phosphatases

In summary, the research community has accumulated a large amount of evidence to support that phosphorylation and dephosphorylation of mitochondrial proteins influence mitochondrial function. However, the strength of the evidence on mitochondrial localization and their activities of the reported kinases and phosphatases vary greatly, and the detailed mechanisms on how these kinases/phosphatases translocate to mitochondria and the physiological and pathological roles related to their mitochondrial localization are still poorly understood. Due to increasing evidence supporting the important functions of mitochondrial kinases/phosphatases, mitochondrial biology is due for more intense exploration in this area.

Abbreviations

ER:

endoplasmic reticulum

AD:

Alzheimer’s disease

CML:

chronic myeloid leukemia

BCR:

breakpoint cluster region

AML:

acute promyelocytic leukemia

PKA:

protein kinase A

AKAP121:

a kinase anchor proteins 121

PTPD1:

protein tyrosine phosphatase D1

ETC:

mitochondrial electron transport chain

EGFRs:

epidermal growth factor receptors

CoxII:

cytochrome c oxidase subunit II

3-MA:

3-methyladenine

SH:

secondary hyper-parathyroidism

Akt:

protein kinase B

LIF:

leukemia inhibitory factor

MPTP:

mitochondrial permeability transition pore

JNK:

c-Jun N-terminal kinase

ERK1/2:

extracellular signal-regulated kinase 1 and 2

p38 MAPKs:

p38 mitogen-activated protein kinases

GSK:

glycogen synthase kinase ANT: adenine nucleotide translocator

HD:

Huntington’s disease

Drp1:

dynamin-related protein 1

CREB:

cAMP response element-binding protein

DS:

down syndrome

Hsps:

heat shock proteins

PKC:

protein kinase C

mitoKATP :

mitochondrial ATP-sensitive K + channel

PINK1:

PTEN-induced kinase 1

PD:

Parkinson’s disease

PARL:

presenilin-associated rhomboid-like serine protease

MKPs:

MAP Kinase Phosphatases

NSCLC:

non-small-cell lung cancer

RA:

rheumatoid arthritis

Shp2:

Src homology 2 domain-containing phosphatase 2

NS:

Noonan syndrome

LS:

Leopard syndrome

JMML:

juvenile myelomonocytic leukemia

MPD:

myeloproliferative disorder

PTP:

protein tyrosine phosphases

PTEN:

phosphatase and tensin homolog deleted on chromosome 10

I/R:

ischemia/reperfusion

TOM:

translocase of the outer membrane of mitochondria

TIM:

translocase of inner mitochondrial membrane

MIA:

mitochondrial intermembrane space assembly machinery

PAM:

presequence translocase-associated motor

SAM:

sorting and assembly machinery

References

  1. Hantschel O, Superti-Furga G. Regulation of the c-Abl and Bcr-Abl tyrosine kinases. Nat Rev Mol Cell Biol. 2004;5(1):33–44.

    Article  CAS  PubMed  Google Scholar 

  2. Kumar S, Bharti A, Mishra NC, Raina D, Kharbanda S, Saxena S, Kufe D. Targeting of the c-Abl tyrosine kinase to mitochondria in the necrotic cell death response to oxidative stress. J Biochem. 2001;276(20):17281–5.

    CAS  Google Scholar 

  3. Ito Y, Pandey P, Mishra N, Kumar S, Narula N, Kharbanda S, Saxena S, Kufe D. Targeting of the c-Abl tyrosine kinase to mitochondria in endoplasmic reticulum stress-induced apoptosis. Mol Cell Biol. 2001;21(18):6233–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Qi X, Mochly-Rosen D. The PKCdelta-Abl complex communicates ER stress to the mitochondria - an essential step in subsequent apoptosis. J Cell Sci. 2008;121(Pt 6):804–13.

    Article  CAS  PubMed  Google Scholar 

  5. Derkinderen P, Scales TM, Hanger DP, Leung KY, Byers HL, Ward MA, Lenz C, Price C, Bird IN, Perera T, et al. Tyrosine 394 is phosphorylated in Alzheimer’s paired helical filament tau and in fetal tau with c-Abl as the candidate tyrosine kinase. J Neurosci. 2005;25(28):6584–93.

    Article  CAS  PubMed  Google Scholar 

  6. Williamson R, Scales T, Clark BR, Gibb G, Reynolds CH, Kellie S, Bird IN, Varndell IM, Sheppard PW, Everall I, et al. Rapid tyrosine phosphorylation of neuronal proteins including tau and focal adhesion kinase in response to amyloid-beta peptide exposure: involvement of Src family protein kinases. J Neurosci. 2002;22(1):10–20.

    CAS  PubMed  Google Scholar 

  7. Alvarez AR, Sandoval PC, Leal NR, Castro PU, Kosik KS. Activation of the neuronal c-Abl tyrosine kinase by amyloid-beta-peptide and reactive oxygen species. Neurobiol Dis. 2004;17(2):326–36.

    Article  CAS  PubMed  Google Scholar 

  8. Ben-Neriah Y, Daley GQ, Mes-Masson AM, Witte ON, Baltimore D. The chronic myelogenous leukemia-specific P210 protein is the product of the bcr/abl hybrid gene. Science. 1986;233(4760):212–4.

    Article  CAS  PubMed  Google Scholar 

  9. Holtz MS, Slovak ML, Zhang F, Sawyers CL, Forman SJ, Bhatia R. Imatinib mesylate (STI571) inhibits growth of primitive malignant progenitors in chronic myelogenous leukemia through reversal of abnormally increased proliferation. Blood. 2002;99(10):3792–800.

    Article  CAS  PubMed  Google Scholar 

  10. Salvi M, Brunati AM, Bordin L, La Rocca N, Clari G, Toninello A. Characterization and location of Src-dependent tyrosine phosphorylation in rat brain mitochondria. Biochim Biophys Acta. 2002;1589(2):181–95.

    Article  CAS  PubMed  Google Scholar 

  11. Miyazaki T, Neff L, Tanaka S, Horne WC, Baron R. Regulation of cytochrome c oxidase activity by c-Src in osteoclasts. J Cell Biol. 2003;160(5):709–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Livigni A, Scorziello A, Agnese S, Adornetto A, Carlucci A, Garbi C, Castaldo I, Annunziato L, Avvedimento EV, Feliciello A. Mitochondrial AKAP121 links cAMP and src signaling to oxidative metabolism. Mol Biol Cell. 2006;17(1):263–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Cardone L, Carlucci A, Affaitati A, Livigni A, DeCristofaro T, Garbi C, Varrone S, Ullrich A, Gottesman ME, Avvedimento EV, et al. Mitochondrial AKAP121 binds and targets protein tyrosine phosphatase D1, a novel positive regulator of src signaling. Mol Cell Biol. 2004;24(11):4613–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Sardanelli AM, Signorile A, Nuzzi R, Rasmo DD, Technikova-Dobrova Z, Drahota Z, Occhiello A, Pica A, Papa S. Occurrence of A-kinase anchor protein and associated cAMP-dependent protein kinase in the inner compartment of mammalian mitochondria. FEBS Lett. 2006;580(24):5690–6.

    Article  CAS  PubMed  Google Scholar 

  15. Itoh S, Lemay S, Osawa M, Che W, Duan Y, Tompkins A, Brookes PS, Sheu SS, Abe J. Mitochondrial Dok-4 recruits Src kinase and regulates NF-kappaB activation in endothelial cells. J Biochem. 2005;280(28):26383–96.

    CAS  Google Scholar 

  16. Arachiche A, Augereau O, Decossas M, Pertuiset C, Gontier E, Letellier T, Dachary-Prigent J. Localization of PTP-1B, SHP-2, and Src exclusively in rat brain mitochondria and functional consequences. J Biochem. 2008;283(36):24406–11.

    CAS  Google Scholar 

  17. Hebert-Chatelain E, Jose C, Cortes NG, Dupuy JW, Rocher C, Dachary-Prigent J, Letellier T. Preservation of NADH ubiquinone-oxidoreductase activity by Src kinase-mediated phosphorylation of NDUFB10. Biochim Biophys Acta. 2012;1817(5):718–25.

    Article  CAS  PubMed  Google Scholar 

  18. Ogura M, Yamaki J, Homma MK, Homma Y. Mitochondrial c-Src regulates cell survival through phosphorylation of respiratory chain components. Biochem J. 2012;447(2):281–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Ishizawar R, Parsons SJ. c-Src and cooperating partners in human cancer. Cancer Cell. 2004;6(3):209–14.

    Article  CAS  PubMed  Google Scholar 

  20. Lluis JM, Buricchi F, Chiarugi P, Morales A, Fernandez-Checa JC. Dual role of mitochondrial reactive oxygen species in hypoxia signaling: activation of nuclear factor-{kappa}B via c-SRC and oxidant-dependent cell death. Cancer Res. 2007;67(15):7368–77.

    Article  CAS  PubMed  Google Scholar 

  21. Tejedo JR, Ramirez R, Cahuana GM, Rincon P, Sobrino F, Bedoya FJ. Evidence for involvement of c-Src in the anti-apoptotic action of nitric oxide in serum-deprived RINm5F cells. Cell Signal. 2001;13(11):809–17.

    Article  CAS  PubMed  Google Scholar 

  22. Myoui A, Nishimura R, Williams PJ, Hiraga T, Tamura D, Michigami T, Mundy GR, Yoneda T. C-SRC tyrosine kinase activity is associated with tumor colonization in bone and lung in an animal model of human breast cancer metastasis. Cancer Res. 2003;63(16):5028–33.

    CAS  PubMed  Google Scholar 

  23. Acin-Perez R, Carrascoso I, Baixauli F, Roche-Molina M, Latorre-Pellicer A, Fernandez-Silva P, Mittelbrunn M, Sanchez-Madrid F, Perez-Martos A, Lowell CA, et al. ROS-triggered phosphorylation of complex II by Fgr kinase regulates cellular adaptation to fuel use. Cell Metab. 2014;19(6):1020–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Hibbs ML, Tarlinton DM, Armes J, Grail D, Hodgson G, Maglitto R, Stacker SA, Dunn AR. Multiple defects in the immune system of Lyn-deficient mice, culminating in autoimmune disease. Cell. 1995;83(2):301–11.

    Article  CAS  PubMed  Google Scholar 

  25. Oracki SA, Tsantikos E, Quilici C, Light A, Schmidt T, Lew AM, Martin JE, Smith KG, Hibbs ML, Tarlinton DM. CTLA4Ig alters the course of autoimmune disease development in Lyn-/- mice. J Immunol. 2010;184(2):757–63.

    Article  CAS  PubMed  Google Scholar 

  26. Roginskaya V, Zuo S, Caudell E, Nambudiri G, Kraker AJ, Corey SJ. Therapeutic targeting of Src-kinase Lyn in myeloid leukemic cell growth. Leukemia. 1999;13(6):855–61.

    Article  CAS  PubMed  Google Scholar 

  27. Donato NJ, Wu JY, Stapley J, Gallick G, Lin H, Arlinghaus R, Talpaz M. BCR-ABL independence and LYN kinase overexpression in chronic myelogenous leukemia cells selected for resistance to STI571. Blood. 2003;101(2):690–8.

    Article  CAS  PubMed  Google Scholar 

  28. Contri A, Brunati AM, Trentin L, Cabrelle A, Miorin M, Cesaro L, Pinna LA, Zambello R, Semenzato G, Donella-Deana A. Chronic lymphocytic leukemia B cells contain anomalous Lyn tyrosine kinase, a putative contribution to defective apoptosis. J Clin Invest. 2005;115(2):369–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Goldenberg-Furmanov M, Stein I, Pikarsky E, Rubin H, Kasem S, Wygoda M, Weinstein I, Reuveni H, Ben-Sasson SA. Lyn is a target gene for prostate cancer: sequence-based inhibition induces regression of human tumor xenografts. Cancer Res. 2004;64(3):1058–66.

    Article  CAS  PubMed  Google Scholar 

  30. Choi YL, Bocanegra M, Kwon MJ, Shin YK, Nam SJ, Yang JH, Kao J, Godwin AK, Pollack JR. LYN is a mediator of epithelial-mesenchymal transition and a target of dasatinib in breast cancer. Cancer Res. 2010;70(6):2296–306.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Stettner MR, Wang W, Nabors LB, Bharara S, Flynn DC, Grammer JR, Gillespie GY, Gladson CL. Lyn kinase activity is the predominant cellular SRC kinase activity in glioblastoma tumor cells. Cancer Res. 2005;65(13):5535–43.

    Article  CAS  PubMed  Google Scholar 

  32. Gringeri E, Carraro A, Tibaldi E, D’Amico FE, Mancon M, Toninello A, Pagano MA, Vio C, Cillo U, Brunati AM. Lyn-mediated mitochondrial tyrosine phosphorylation is required to preserve mitochondrial integrity in early liver regeneration. Biochem J. 2010;425(2):401–12.

    Article  CAS  Google Scholar 

  33. Boerner JL, Demory ML, Silva C, Parsons SJ. Phosphorylation of Y845 on the epidermal growth factor receptor mediates binding to the mitochondrial protein cytochrome c oxidase subunit II. Mol Cell Biol. 2004;24(16):7059–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Demory ML, Boerner JL, Davidson R, Faust W, Miyake T, Lee I, Huttemann M, Douglas R, Haddad G, Parsons SJ. Epidermal growth factor receptor translocation to the mitochondria: regulation and effect. J Biochem. 2009;284(52):36592–604.

    CAS  Google Scholar 

  35. Cao X, Zhu H, Ali-Osman F, Lo HW. EGFR and EGFRvIII undergo stress- and EGFR kinase inhibitor-induced mitochondrial translocalization: a potential mechanism of EGFR-driven antagonism of apoptosis. Mol Cancer. 2011;10:26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Dreier A, Barth S, Goswami A, Weis J. Cetuximab induces mitochondrial translocalization of EGFRvIII, but not EGFR: involvement of mitochondria in tumor drug resistance? Tumour Biol. 2012;33(1):85–94.

    Article  CAS  PubMed  Google Scholar 

  37. Dasari VR, Velpula KK, Alapati K, Gujrati M, Tsung AJ. Cord blood stem cells inhibit epidermal growth factor receptor translocation to mitochondria in glioblastoma. PLoS One. 2012;7(2):e31884.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Yue X, Song W, Zhang W, Chen L, Xi Z, Xin Z, Jiang X. Mitochondrially localized EGFR is subjected to autophagic regulation and implicated in cell survival. Autophagy. 2008;4(5):641–9.

    Article  CAS  PubMed  Google Scholar 

  39. Yao Y, Wang G, Li Z, Yan B, Guo Y, Jiang X, Xi Z. Mitochondrially localized EGFR is independent of its endocytosis and associates with cell viability. Acta Biochim Biophys Sin. 2010;42(11):763–70.

    Article  CAS  PubMed  Google Scholar 

  40. Ding Y, Liu Z, Desai S, Zhao Y, Liu H, Pannell LK, Yi H, Wright ER, Owen LB, Dean-Colomb W, et al. Receptor tyrosine kinase ErbB2 translocates into mitochondria and regulates cellular metabolism. Nat Commun. 2012;3:1271.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Holbro T, Hynes NE. ErbB receptors: directing key signaling networks throughout life. Ann Rev Pharmacol Toxicol. 2004;44:195–217.

    Article  CAS  Google Scholar 

  42. Sasaki K, Sato M, Umezawa Y. Fluorescent indicators for Akt/protein kinase B and dynamics of Akt activity visualized in living cells. J Biochem. 2003;278(33):30945–51.

    CAS  Google Scholar 

  43. Bijur GN, Jope RS. Rapid accumulation of Akt in mitochondria following phosphatidylinositol 3-kinase activation. J Neurochem. 2003;87(6):1427–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Miyamoto S, Murphy AN, Brown JH. Akt mediates mitochondrial protection in cardiomyocytes through phosphorylation of mitochondrial hexokinase-II. Cell Death Differ. 2008;15(3):521–9.

    Article  CAS  PubMed  Google Scholar 

  45. Yang JY, Deng W, Chen Y, Fan W, Baldwin KM, Jope RS, Wallace DC, Wang PH. Impaired translocation and activation of mitochondrial Akt1 mitigated mitochondrial oxidative phosphorylation complex V activity in diabetic myocardium. J Mol Cell Cardiol. 2013;59:167–75.

    Article  CAS  PubMed  Google Scholar 

  46. Datta SR, Dudek H, Tao X, Masters S, Fu H, Gotoh Y, Greenberg ME. Akt phosphorylation of BAD couples survival signals to the cell-intrinsic death machinery. Cell. 1997;91(2):231–41.

    Article  CAS  PubMed  Google Scholar 

  47. Majewski M, Nieborowska-Skorska M, Salomoni P, Slupianek A, Reiss K, Trotta R, Calabretta B, Skorski T. Activation of mitochondrial Raf-1 is involved in the antiapoptotic effects of Akt. Cancer Res. 1999;59(12):2815–9.

    CAS  PubMed  Google Scholar 

  48. Yang JY, Yeh HY, Lin K, Wang PH. Insulin stimulates Akt translocation to mitochondria: implications on dysregulation of mitochondrial oxidative phosphorylation in diabetic myocardium. J Mol Cell Cardiol. 2009;46(6):919–26.

    Article  CAS  PubMed  Google Scholar 

  49. Barksdale KA, Bijur GN. The basal flux of Akt in the mitochondria is mediated by heat shock protein 90. J Neurochem. 2009;108(5):1289–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Ito Y, Mishra NC, Yoshida K, Kharbanda S, Saxena S, Kufe D. Mitochondrial targeting of JNK/SAPK in the phorbol ester response of myeloid leukemia cells. Cell Death Differ. 2001;8(8):794–800.

    Article  CAS  PubMed  Google Scholar 

  51. Baines CP, Zhang J, Wang GW, Zheng YT, Xiu JX, Cardwell EM, Bolli R, Ping P. Mitochondrial PKCepsilon and MAPK form signaling modules in the murine heart: enhanced mitochondrial PKCepsilon-MAPK interactions and differential MAPK activation in PKCepsilon-induced cardioprotection. Circ Res. 2002;90(4):390–7.

    Article  CAS  PubMed  Google Scholar 

  52. Chauhan D, Li G, Hideshima T, Podar K, Mitsiades C, Mitsiades N, Munshi N, Kharbanda S, Anderson KC. JNK-dependent release of mitochondrial protein, Smac, during apoptosis in multiple myeloma (MM) cells. J Biochem. 2003;278(20):17593–6.

    CAS  Google Scholar 

  53. Rumora L, Lovric J, Sairam MR, Maysinger D. Impairments of heat shock protein expression and MAPK translocation in the central nervous system of follitropin receptor knockout mice. Exp Gerontol. 2007;42(7):619–28.

    Article  CAS  PubMed  Google Scholar 

  54. Hanawa N, Shinohara M, Saberi B, Gaarde WA, Han D, Kaplowitz N. Role of JNK translocation to mitochondria leading to inhibition of mitochondria bioenergetics in acetaminophen-induced liver injury. J Biochem. 2008;283(20):13565–77.

    CAS  Google Scholar 

  55. Zhou Q, Lam PY, Han D, Cadenas E. c-Jun N-terminal kinase regulates mitochondrial bioenergetics by modulating pyruvate dehydrogenase activity in primary cortical neurons. J Neurochem. 2008;104(2):325–35.

    CAS  PubMed  Google Scholar 

  56. Nijboer CH, Bonestroo HJ, Zijlstra J, Kavelaars A, Heijnen CJ. Mitochondrial JNK phosphorylation as a novel therapeutic target to inhibit neuroinflammation and apoptosis after neonatal ischemic brain damage. Neurobiol Dis. 2013;54:432–44.

    Article  CAS  PubMed  Google Scholar 

  57. Chambers JW, Howard S, LoGrasso PV. Blocking c-Jun N-terminal kinase (JNK) translocation to the mitochondria prevents 6-hydroxydopamine-induced toxicity in vitro and in vivo. J Biochem. 2013;288(2):1079–87.

    CAS  Google Scholar 

  58. Bhakar AL, Howell JL, Paul CE, Salehi AH, Becker EB, Said F, Bonni A, Barker PA. Apoptosis induced by p75NTR overexpression requires Jun kinase-dependent phosphorylation of Bad. J Neurosci. 2003;23(36):11373–81.

    CAS  PubMed  Google Scholar 

  59. Yu C, Minemoto Y, Zhang J, Liu J, Tang F, Bui TN, Xiang J, Lin A. JNK suppresses apoptosis via phosphorylation of the proapoptotic Bcl-2 family protein BAD. Mol Cell. 2004;13(3):329–40.

    Article  CAS  PubMed  Google Scholar 

  60. Deng X, Xiao L, Lang W, Gao F, Ruvolo P, May WS Jr. Novel role for JNK as a stress-activated Bcl2 kinase. J Biochem. 2001;276(26):23681–8.

    CAS  Google Scholar 

  61. Nowak G, Clifton GL, Godwin ML, Bakajsova D. Activation of ERK1/2 pathway mediates oxidant-induced decreases in mitochondrial function in renal cells. Am J Physiol Renal Physiol. 2006;291(4):F840–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Monick MM, Powers LS, Barrett CW, Hinde S, Ashare A, Groskreutz DJ, Nyunoya T, Coleman M, Spitz DR, Hunninghake GW. Constitutive ERK MAPK activity regulates macrophage ATP production and mitochondrial integrity. J Immunol. 2008;180(11):7485–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Dagda RK, Zhu J, Kulich SM, Chu CT. Mitochondrially localized ERK2 regulates mitophagy and autophagic cell stress: implications for Parkinson’s disease. Autophagy. 2008;4(6):770–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Zhuang S, Kinsey GR, Yan Y, Han J, Schnellmann RG. Extracellular signal-regulated kinase activation mediates mitochondrial dysfunction and necrosis induced by hydrogen peroxide in renal proximal tubular cells. J Pharmacol Exp Therap. 2008;325(3):732–40.

    Article  CAS  Google Scholar 

  65. Hernandez-Resendiz S, Zazueta C. PHO-ERK1/2 interaction with mitochondria regulates the permeability transition pore in cardioprotective signaling. Life Sci. 2014;108(1):13–21.

    Article  CAS  PubMed  Google Scholar 

  66. Duarte A, Castillo AF, Podesta EJ, Poderoso C. Mitochondrial fusion and ERK activity regulate steroidogenic acute regulatory protein localization in mitochondria. PLoS One. 2014;9(6):e100387.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Ballard-Croft C, Kristo G, Yoshimura Y, Reid E, Keith BJ, Mentzer RM Jr, Lasley RD. Acute adenosine preconditioning is mediated by p38 MAPK activation in discrete subcellular compartments. Am J Physiol Heart Circ Physiol. 2005;288(3):H1359–66.

    Article  CAS  PubMed  Google Scholar 

  68. Galli S, Antico Arciuch VG, Poderoso C, Converso DP, Zhou Q, Bal de Kier Joffe E, Cadenas E, Boczkowski J, Carreras MC, Poderoso JJ. Tumor cell phenotype is sustained by selective MAPK oxidation in mitochondria. PLoS One. 2008;3(6):2379.

    Article  CAS  Google Scholar 

  69. Hoshi M, Sato M, Kondo S, Takashima A, Noguchi K, Takahashi M, Ishiguro K, Imahori K. Different localization of tau protein kinase I/glycogen synthase kinase-3 beta from glycogen synthase kinase-3 alpha in cerebellum mitochondria. J Biochem. 1995;118(4):683–5.

    CAS  PubMed  Google Scholar 

  70. Bijur GN, Jope RS. Glycogen synthase kinase-3 beta is highly activated in nuclei and mitochondria. Neuro Rep. 2003;14(18):2415–9.

    CAS  Google Scholar 

  71. Nishihara M, Miura T, Miki T, Tanno M, Yano T, Naitoh K, Ohori K, Hotta H, Terashima Y, Shimamoto K. Modulation of the mitochondrial permeability transition pore complex in GSK-3beta-mediated myocardial protection. J Mol Cell Cardiol. 2007;43(5):564–70.

    Article  CAS  PubMed  Google Scholar 

  72. Xi J, Wang H, Mueller RA, Norfleet EA, Xu Z. Mechanism for resveratrol-induced cardioprotection against reperfusion injury involves glycogen synthase kinase 3beta and mitochondrial permeability transition pore. Eur J Pharmacol. 2009;604(1–3):111–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Rodriguez-Sinovas A, Boengler K, Cabestrero A, Gres P, Morente M, Ruiz-Meana M, Konietzka I, Miro E, Totzeck A, Heusch G, et al. Translocation of connexin 43 to the inner mitochondrial membrane of cardiomyocytes through the heat shock protein 90-dependent TOM pathway and its importance for cardioprotection. Circ Res. 2006;99(1):93–101.

    Article  CAS  PubMed  Google Scholar 

  74. Hoshi M, Takashima A, Noguchi K, Murayama M, Sato M, Kondo S, Saitoh Y, Ishiguro K, Hoshino T, Imahori K. Regulation of mitochondrial pyruvate dehydrogenase activity by tau protein kinase I/glycogen synthase kinase 3beta in brain. Proc Nat Acad Sci USA. 1996;93(7):2719–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Juhaszova M, Zorov DB, Kim SH, Pepe S, Fu Q, Fishbein KW, Ziman BD, Wang S, Ytrehus K, Antos CL, et al. Glycogen synthase kinase-3beta mediates convergence of protection signaling to inhibit the mitochondrial permeability transition pore. J Clin Invest. 2004;113(11):1535–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Hanger DP, Hughes K, Woodgett JR, Brion JP, Anderton BH. Glycogen synthase kinase-3 induces Alzheimer’s disease-like phosphorylation of tau: generation of paired helical filament epitopes and neuronal localisation of the kinase. Neurosci Lett. 1992;147(1):58–62.

    Article  CAS  PubMed  Google Scholar 

  77. Mulot SF, Hughes K, Woodgett JR, Anderton BH, Hanger DP. PHF-tau from Alzheimer’s brain comprises four species on SDS-PAGE which can be mimicked by in vitro phosphorylation of human brain tau by glycogen synthase kinase-3 beta. FEBS Lett. 1994;349(3):359–64.

    Article  CAS  PubMed  Google Scholar 

  78. Oliveira JM, Henriques AG, Martins F, Rebelo S, da Cruz E Silva OA. Amyloid-beta modulates both AβPP and tau phosphorylation. J Alzheimer Dis. 2015;45(2):495–507.

    CAS  Google Scholar 

  79. Kwok JB, Hallupp M, Loy CT, Chan DK, Woo J, Mellick GD, Buchanan DD, Silburn PA, Halliday GM, Schofield PR. GSK3B polymorphisms alter transcription and splicing in Parkinson’s disease. Ann Neurol. 2005;58(6):829–39.

    Article  CAS  PubMed  Google Scholar 

  80. Ma C, Wang J, Gao Y, Gao TW, Chen G, Bower KA, Odetallah M, Ding M, Ke Z, Luo J. The role of glycogen synthase kinase 3beta in the transformation of epidermal cells. Cancer Res. 2007;67(16):7756–64.

    Article  CAS  PubMed  Google Scholar 

  81. Farago M, Dominguez I, Landesman-Bollag E, Xu X, Rosner A, Cardiff RD, Seldin DC. Kinase-inactive glycogen synthase kinase 3beta promotes Wnt signaling and mammary tumorigenesis. Cancer Res. 2005;65(13):5792–801.

    Article  CAS  PubMed  Google Scholar 

  82. Vincent T, Kukalev A, Andang M, Pettersson R, Percipalle P. The glycogen synthase kinase (GSK) 3beta represses RNA polymerase I transcription. Oncogene. 2008;27(39):5254–9.

    Article  CAS  PubMed  Google Scholar 

  83. Shakoori A, Mai W, Miyashita K, Yasumoto K, Takahashi Y, Ooi A, Kawakami K, Minamoto T. Inhibition of GSK-3 beta activity attenuates proliferation of human colon cancer cells in rodents. Cancer Sci. 2007;98(9):1388–93.

    Article  CAS  PubMed  Google Scholar 

  84. Ougolkov AV, Fernandez-Zapico ME, Savoy DN, Urrutia RA, Billadeau DD. Glycogen synthase kinase-3beta participates in nuclear factor kappaB-mediated gene transcription and cell survival in pancreatic cancer cells. Cancer Res. 2005;65(6):2076–81.

    Article  CAS  PubMed  Google Scholar 

  85. Papa S, Sardanelli AM, Scacco S, Technikova-Dobrova Z. cAMP-dependent protein kinase and phosphoproteins in mammalian mitochondria. An extension of the cAMP-mediated intracellular signal transduction. FEBS Lett. 1999;444(2):245–9.

    Article  CAS  PubMed  Google Scholar 

  86. Kleitke B, Sydow H, Wollenberger A. Evidence for cyclic AMP-dependent protein kinase activity in isolated guinea pig and rat liver mitochondria. Acta Biologica et medica Germanica. 1976;35(3–4):K9–17.

    CAS  PubMed  Google Scholar 

  87. Schwoch G, Trinczek B, Bode C. Localization of catalytic and regulatory subunits of cyclic AMP-dependent protein kinases in mitochondria from various rat tissues. Biochem J. 1990;270(1):181–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Verdanis A. Protein kinase activity at the inner membrane of mammalian mitochondria. J Biochem. 1977;252(3):807–13.

    CAS  Google Scholar 

  89. Ma MP, Thomson M. Protein Kinase A Subunit alpha Catalytic and A Kinase Anchoring Protein 79 in Human Placental Mitochondria. Open Biochem J. 2012;6:23–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Sardanelli AM, Technikova-Dobrova Z, Speranza F, Mazzocca A, Scacco S, Papa S. Topology of the mitochondrial cAMP-dependent protein kinase and its substrates. FEBS Lett. 1996;396(2–3):276–8.

    Article  CAS  PubMed  Google Scholar 

  91. Technikova-Dobrova Z, Sardanelli AM, Papa S. Phosphorylation of mitochondrial proteins in bovine heart. Characterization of kinases and substrates. FEBS Lett. 1993;322(1):51–5.

    Article  CAS  PubMed  Google Scholar 

  92. Technikova-Dobrova Z, Sardanelli AM, Stanca MR, Papa S. cAMP-dependent protein phosphorylation in mitochondria of bovine heart. FEBS Lett. 1994;350(2–3):187–91.

    Article  CAS  PubMed  Google Scholar 

  93. Gomez-Chang E, Espinosa-Garcia MT, Olvera-Sanchez S, Flores-Herrera O, Martinez F. PKA tightly bound to human placental mitochondria participates in steroidogenesis and is not modified by cAMP. Placenta. 2014;35(9):748–62.

    Article  CAS  PubMed  Google Scholar 

  94. Feliciello A, Gottesman ME, Avvedimento EV. The biological functions of A-kinase anchor proteins. J Mol Biol. 2001;308(2):99–114.

    Article  CAS  PubMed  Google Scholar 

  95. Tasken K, Aandahl EM. Localized effects of cAMP mediated by distinct routes of protein kinase A. Physiol Rev. 2004;84(1):137–67.

    Article  CAS  PubMed  Google Scholar 

  96. Harada H, Becknell B, Wilm M, Mann M, Huang LJ, Taylor SS, Scott JD, Korsmeyer SJ. Phosphorylation and inactivation of BAD by mitochondria-anchored protein kinase A. Mol Cell. 1999;3(4):413–22.

    Article  CAS  PubMed  Google Scholar 

  97. Affaitati A, Cardone L, de Cristofaro T, Carlucci A, Ginsberg MD, Varrone S, Gottesman ME, Avvedimento EV, Feliciello A. Essential role of A-kinase anchor protein 121 for cAMP signaling to mitochondria. J Biochem. 2003;278(6):4286–94.

    CAS  Google Scholar 

  98. Merrill RA, Strack S. Mitochondria: a kinase anchoring protein 1, a signaling platform for mitochondrial form and function. Int J Biochem Cell Biol. 2014;48:92–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Cribbs JT, Strack S. Reversible phosphorylation of Drp1 by cyclic AMP-dependent protein kinase and calcineurin regulates mitochondrial fission and cell death. EMBO Rep. 2007;8(10):939–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Dickey AS, Strack S. PKA/AKAP1 and PP2A/Bbeta2 regulate neuronal morphogenesis via Drp1 phosphorylation and mitochondrial bioenergetics. J Neurosci. 2011;31(44):15716–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Lee J, Kim CH, Simon DK, Aminova LR, Andreyev AY, Kushnareva YE, Murphy AN, Lonze BE, Kim KS, Ginty DD, et al. Mitochondrial cyclic AMP response element-binding protein (CREB) mediates mitochondrial gene expression and neuronal survival. J Biochem. 2005;280(49):40398–401.

    CAS  Google Scholar 

  102. Ryu H, Lee J, Impey S, Ratan RR, Ferrante RJ. Antioxidants modulate mitochondrial PKA and increase CREB binding to D-loop DNA of the mitochondrial genome in neurons. Proc Nat Acad Sci USA. 2005;102(39):13915–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Beall A, Bagwell D, Woodrum D, Stoming TA, Kato K, Suzuki A, Rasmussen H, Brophy CM. The small heat shock-related protein, HSP20, is phosphorylated on serine 16 during cyclic nucleotide-dependent relaxation. J Biochem. 1999;274(16):11344–51.

    CAS  Google Scholar 

  104. Edwards HV, Scott JD, Baillie GS. PKA phosphorylation of the small heat-shock protein Hsp20 enhances its cardioprotective effects. Biochem Soc Trans. 2012;40(1):210–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Ohnuma Y, Miura T, Miki T, Tanno M, Kuno A, Tsuchida A, Shimamoto K. Opening of mitochondrial K(ATP) channel occurs downstream of PKC-epsilon activation in the mechanism of preconditioning. Am J Physiol Heart Circ Physiol. 2002;283(1):H440–7.

    Article  CAS  PubMed  Google Scholar 

  106. Ogbi M, Johnson JA. Protein kinase Cepsilon interacts with cytochrome c oxidase subunit IV and enhances cytochrome c oxidase activity in neonatal cardiac myocyte preconditioning. Biochem J. 2006;393(Pt 1):191–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Ping P, Song C, Zhang J, Guo Y, Cao X, Li RC, Wu W, Vondriska TM, Pass JM, Tang XL, et al. Formation of protein kinase C(epsilon)-Lck signaling modules confers cardioprotection. J Clin Invest. 2002;109(4):499–507.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Jaburek M, Costa AD, Burton JR, Costa CL, Garlid KD. Mitochondrial PKC epsilon and mitochondrial ATP-sensitive K + channel copurify and coreconstitute to form a functioning signaling module in proteoliposomes. Circ Res. 2006;99(8):878–83.

    Article  CAS  PubMed  Google Scholar 

  109. Waza AA, Andrabi K, Hussain MU. Protein kinase C (PKC) mediated interaction between conexin43 (Cx43) and K(+)(ATP) channel subunit (Kir6.1) in cardiomyocyte mitochondria: implications in cytoprotection against hypoxia induced cell apoptosis. Cell Signal. 2014;26(9):1909–17.

    Article  CAS  PubMed  Google Scholar 

  110. Hamasaki Y, Shinohara O, Ishida H, Hayashi Y, Nakazawa H. Decreased protein kinase C-epsilon expression in hypertrophied cardiac ventricles induced by triiodothyronine treatment in the rat. Life Sci. 2000;67(15):1859–68.

    Article  CAS  PubMed  Google Scholar 

  111. Inagaki K, Iwanaga Y, Sarai N, Onozawa Y, Takenaka H, Mochly-Rosen D, Kihara Y. Tissue angiotensin II during progression or ventricular hypertrophy to heart failure in hypertensive rats; differential effects on PKC epsilon and PKC beta. J Mol Cell Cardiol. 2002;34(10):1377–85.

    Article  CAS  PubMed  Google Scholar 

  112. Chen L, Hahn H, Wu G, Chen CH, Liron T, Schechtman D, Cavallaro G, Banci L, Guo Y, Bolli R, et al. Opposing cardioprotective actions and parallel hypertrophic effects of delta PKC and epsilon PKC. Proc Nat Acad Sci USA. 2001;98(20):11114–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Choi DS, Wang D, Yu GQ, Zhu G, Kharazia VN, Paredes JP, Chang WS, Deitchman JK, Mucke L, Messing RO. PKCepsilon increases endothelin converting enzyme activity and reduces amyloid plaque pathology in transgenic mice. Proc Nat Acad Sci USA. 2006;103(21):8215–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Majumder PK, Pandey P, Sun X, Cheng K, Datta R, Saxena S, Kharbanda S, Kufe D. Mitochondrial translocation of protein kinase C delta in phorbol ester-induced cytochrome c release and apoptosis. J Biochem. 2000;275(29):21793–6.

    CAS  Google Scholar 

  115. Majumder PK, Mishra NC, Sun X, Bharti A, Kharbanda S, Saxena S, Kufe D. Targeting of protein kinase C delta to mitochondria in the oxidative stress response. Cell Grow Differ. 2001;12(9):465–70.

    CAS  Google Scholar 

  116. Sumitomo M, Ohba M, Asakuma J, Asano T, Kuroki T, Hayakawa M. Protein kinase Cdelta amplifies ceramide formation via mitochondrial signaling in prostate cancer cells. J Clin Invest. 2002;109(6):827–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Li L, Lorenzo PS, Bogi K, Blumberg PM, Yuspa SH. Protein kinase Cdelta targets mitochondria, alters mitochondrial membrane potential, and induces apoptosis in normal and neoplastic keratinocytes when overexpressed by an adenoviral vector. Mol Cell Biol. 1999;19(12):8547–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Valente EM, Abou-Sleiman PM, Caputo V, Muqit MM, Harvey K, Gispert S, Ali Z, Del Turco D, Bentivoglio AR, Healy DG, et al. Hereditary early-onset Parkinson’s disease caused by mutations in PINK1. Science. 2004;304(5674):1158–60.

    Article  CAS  PubMed  Google Scholar 

  119. Weihofen A, Thomas KJ, Ostaszewski BL, Cookson MR, Selkoe DJ. Pink1 forms a multiprotein complex with Miro and Milton, linking Pink1 function to mitochondrial trafficking. Biochemistry. 2009;48(9):2045–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Jin SM, Lazarou M, Wang C, Kane LA, Narendra DP, Youle RJ. Mitochondrial membrane potential regulates PINK1 import and proteolytic destabilization by PARL. J Cell Biol. 2010;191(5):933–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Gasser T. Genetics of Parkinson’s disease. Dial Clin Neurosci. 2004;6(3):295–301.

    Google Scholar 

  122. Chen Y, Sparks M, Bhandari P, Matkovich SJ, Dorn GW. Mitochondrial genome linearization is a causative factor for cardiomyopathy in mice and drosophila. Antioxid Redox Signal. 2013;21(4):1949–59.

    PubMed  Google Scholar 

  123. Wang X, Winter D, Ashrafi G, Schlehe J, Wong YL, Selkoe D, Rice S, Steen J, LaVoie MJ, Schwarz TL. PINK1 and Parkin target Miro for phosphorylation and degradation to arrest mitochondrial motility. Cell. 2011;147(4):893–906.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Dagda RK, Cherra SJ 3rd, Kulich SM, Tandon A, Park D, Chu CT. Loss of PINK1 function promotes mitophagy through effects on oxidative stress and mitochondrial fission. J Biochem. 2009;284(20):13843–55.

    CAS  Google Scholar 

  125. Lutz AK, Exner N, Fett ME, Schlehe JS, Kloos K, Lammermann K, Brunner B, Kurz-Drexler A, Vogel F, Reichert AS, et al. Loss of parkin or PINK1 function increases Drp1-dependent mitochondrial fragmentation. J Biochem. 2009;284(34):22938–51.

    CAS  Google Scholar 

  126. Liu S, Sawada T, Lee S, Yu W, Silverio G, Alapatt P, Millan I, Shen A, Saxton W, Kanao T, et al. Parkinson’s disease-associated kinase PINK1 regulates Miro protein level and axonal transport of mitochondria. PLoS Genet. 2012;8(3):e1002537.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Gautier CA, Kitada T, Shen J. Loss of PINK1 causes mitochondrial functional defects and increased sensitivity to oxidative stress. Proc Nat Acad Sci USA. 2008;105(32):11364–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Morais VA, Verstreken P, Roethig A, Smet J, Snellinx A, Vanbrabant M, Haddad D, Frezza C, Mandemakers W, Vogt-Weisenhorn D, et al. Parkinson’s disease mutations in PINK1 result in decreased Complex I activity and deficient synaptic function. EMBO Mol Med. 2009;1(2):99–111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Vos M, Esposito G, Edirisinghe JN, Vilain S, Haddad DM, Slabbaert JR, Van Meensel S, Schaap O, De Strooper B, Meganathan R, et al. Vitamin K2 is a mitochondrial electron carrier that rescues pink1 deficiency. Science. 2012;336(6086):1306–10.

    Article  CAS  PubMed  Google Scholar 

  130. Vincow ES, Merrihew G, Thomas RE, Shulman NJ, Beyer RP, MacCoss MJ, Pallanck LJ. The PINK1-Parkin pathway promotes both mitophagy and selective respiratory chain turnover in vivo. Proc Nat Acad Sci USA. 2013;110(16):6400–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Gehrke S, Wu Z, Klinkenberg M, Sun Y, Auburger G, Guo S, Lu B. PINK1 and Parkin control localized translation of respiratory chain component mRNAs on mitochondria outer membrane. Cell Metab. 2015;21(1):95–108.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Sekine S, Kanamaru Y, Koike M, Nishihara A, Okada M, Kinoshita H, Kamiyama M, Maruyama J, Uchiyama Y, Ishihara N, et al. Rhomboid protease PARL mediates the mitochondrial membrane potential loss-induced cleavage of PGAM5. J Biochem. 2012;287(41):34635–45.

    CAS  Google Scholar 

  133. Moisoi N, Fedele V, Edwards J, Martins LM. Loss of PINK1 enhances neurodegeneration in a mouse model of Parkinson’s disease triggered by mitochondrial stress. Neuropharmacology. 2014;77:350–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Tan EK, Refai FS, Siddique M, Yap K, Ho P, Fook-Chong S, Zhao Y. Clinically reported heterozygous mutations in the PINK1 kinase domain exert a gene dosage effect. Hum Mutat. 2009;30(11):1551–7.

    Article  CAS  PubMed  Google Scholar 

  135. Sim CH, Lio DS, Mok SS, Masters CL, Hill AF, Culvenor JG, Cheng HC. C-terminal truncation and Parkinson’s disease-associated mutations down-regulate the protein serine/threonine kinase activity of PTEN-induced kinase-1. Hum Mol Genet. 2006;15(21):3251–62.

    Article  CAS  PubMed  Google Scholar 

  136. Inzelberg R, Jankovic J. Are Parkinson disease patients protected from some but not all cancers? Neurology. 2007;69(15):1542–50.

    Article  PubMed  Google Scholar 

  137. Siddall HK, Warrell CE, Davidson SM, Mocanu MM, Yellon DM. Mitochondrial PINK1–a novel cardioprotective kinase? Cardiovasc Drug. 2008;22(6):507–8.

    Article  Google Scholar 

  138. Scheele C, Nielsen AR, Walden TB, Sewell DA, Fischer CP, Brogan RJ, Petrovic N, Larsson O, Tesch PA, Wennmalm K, et al. Altered regulation of the PINK1 locus: a link between type 2 diabetes and neurodegeneration? FASEB J. 2007;21(13):3653–65.

    Article  CAS  PubMed  Google Scholar 

  139. Gandhi S, Wood-Kaczmar A, Yao Z, Plun-Favreau H, Deas E, Klupsch K, Downward J, Latchman DS, Tabrizi SJ, Wood NW, et al. PINK1-associated Parkinson’s disease is caused by neuronal vulnerability to calcium-induced cell death. Mol Cell. 2009;33(5):627–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Deas E, Piipari K, Machhada A, Li A, Gutierrez-del-Arroyo A, Withers DJ, Wood NW, Abramov AY. PINK1 deficiency in beta-cells increases basal insulin secretion and improves glucose tolerance in mice. Open Biol. 2014;4:140051.

    Article  PubMed Central  CAS  Google Scholar 

  141. Wu JJ, Zhang L, Bennett AM. The noncatalytic amino terminus of mitogen-activated protein kinase phosphatase 1 directs nuclear targeting and serum response element transcriptional regulation. Mol Cell Biol. 2005;25(11):4792–803.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Rosini P, De Chiara G, Bonini P, Lucibello M, Marcocci ME, Garaci E, Cozzolino F, Torcia M. Nerve growth factor-dependent survival of CESS B cell line is mediated by increased expression and decreased degradation of MAPK phosphatase 1. J Biochem. 2004;279(14):14016–23.

    CAS  Google Scholar 

  143. Wang HY, Cheng Z, Malbon CC. Overexpression of mitogen-activated protein kinase phosphatases MKP1, MKP2 in human breast cancer. Cancer Lett. 2003;191(2):229–37.

    Article  CAS  PubMed  Google Scholar 

  144. Loda M, Capodieci P, Mishra R, Yao H, Corless C, Grigioni W, Wang Y, Magi-Galluzzi C, Stork PJ. Expression of mitogen-activated protein kinase phosphatase-1 in the early phases of human epithelial carcinogenesis. Am J Pathol. 1996;149(5):1553–64.

    CAS  PubMed  PubMed Central  Google Scholar 

  145. Denkert C, Schmitt WD, Berger S, Reles A, Pest S, Siegert A, Lichtenegger W, Dietel M, Hauptmann S. Expression of mitogen-activated protein kinase phosphatase-1 (MKP-1) in primary human ovarian carcinoma. Int J Cancer. 2002;102(5):507–13.

    Article  CAS  PubMed  Google Scholar 

  146. Chattopadhyay S, Machado-Pinilla R, Manguan-Garcia C, Belda-Iniesta C, Moratilla C, Cejas P, Fresno-Vara JA, de Castro-Carpeno J, Casado E, Nistal M, et al. MKP1/CL100 controls tumor growth and sensitivity to cisplatin in non-small-cell lung cancer. Oncogene. 2006;25(23):3335–45.

    Article  CAS  PubMed  Google Scholar 

  147. Moncho-Amor V, de Caceres II, Bandres E, Martinez-Poveda B, Orgaz JL, Sanchez-Perez I, Zazo S, Rovira A, Albanell J, Jimenez B, et al. DUSP1/MKP1 promotes angiogenesis, invasion and metastasis in non-small-cell lung cancer. Oncogene. 2011;30(6):668–78.

    Article  CAS  PubMed  Google Scholar 

  148. Salojin KV, Owusu IB, Millerchip KA, Potter M, Platt KA, Oravecz T. Essential role of MAPK phosphatase-1 in the negative control of innate immune responses. J Immunol. 2006;176(3):1899–907.

    Article  CAS  PubMed  Google Scholar 

  149. Chi H, Barry SP, Roth RJ, Wu JJ, Jones EA, Bennett AM, Flavell RA. Dynamic regulation of pro- and anti-inflammatory cytokines by MAPK phosphatase 1 (MKP-1) in innate immune responses. Proc Nat Acad Sci USA. 2006;103(7):2274–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Zhao Q, Shepherd EG, Manson ME, Nelin LD, Sorokin A, Liu Y. The role of mitogen-activated protein kinase phosphatase-1 in the response of alveolar macrophages to lipopolysaccharide: attenuation of proinflammatory cytokine biosynthesis via feedback control of p38. J Biochem. 2005;280(9):8101–8.

    CAS  Google Scholar 

  151. Salvi M, Stringaro A, Brunati AM, Agostinelli E, Arancia G, Clari G, Toninello A. Tyrosine phosphatase activity in mitochondria: presence of Shp-2 phosphatase in mitochondria. Cell Mol Life Sci. 2004;61(18):2393–404.

    Article  CAS  PubMed  Google Scholar 

  152. Zang QS, Martinez B, Yao X, Maass DL, Ma L, Wolf SE, Minei JP. Sepsis-induced cardiac mitochondrial dysfunction involves altered mitochondrial-localization of tyrosine kinase Src and tyrosine phosphatase SHP2. PLoS One. 2012;7(8):e43424.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Tartaglia M, Mehler EL, Goldberg R, Zampino G, Brunner HG, Kremer H, van der Burgt I, Crosby AH, Ion A, Jeffery S, et al. Mutations in PTPN11, encoding the protein tyrosine phosphatase SHP-2, cause Noonan syndrome. Nature Genet. 2001;29(4):465–8.

    Article  CAS  PubMed  Google Scholar 

  154. Legius E, Schrander-Stumpel C, Schollen E, Pulles-Heintzberger C, Gewillig M, Fryns JP. PTPN11 mutations in LEOPARD syndrome. J Med Genet. 2002;39(8):571–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Tartaglia M, Gelb BD. Noonan syndrome and related disorders: genetics and pathogenesis. Ann Rev Genomics Hum Genet. 2005;6:45–68.

    Article  CAS  Google Scholar 

  156. Kontaridis MI, Swanson KD, David FS, Barford D, Neel BG. PTPN11 (Shp2) mutations in LEOPARD syndrome have dominant negative, not activating, effects. J Biochem. 2006;281(10):6785–92.

    CAS  Google Scholar 

  157. Loh ML, Vattikuti S, Schubbert S, Reynolds MG, Carlson E, Lieuw KH, Cheng JW, Lee CM, Stokoe D, Bonifas JM, et al. Mutations in PTPN11 implicate the SHP-2 phosphatase in leukemogenesis. Blood. 2004;103(6):2325–31.

    Article  CAS  PubMed  Google Scholar 

  158. Emanuel PD, Shannon KM, Castleberry RP. Juvenile myelomonocytic leukemia: molecular understanding and prospects for therapy. Mol Med Today. 1996;2(11):468–75.

    Article  CAS  PubMed  Google Scholar 

  159. Merks JH, Caron HN, Hennekam RC. High incidence of malformation syndromes in a series of 1073 children with cancer. Am J Med Genet Part A. 2005;134A(2):132–43.

    Article  PubMed  Google Scholar 

  160. Ucar C, Calyskan U, Martini S, Heinritz W. Acute myelomonocytic leukemia in a boy with LEOPARD syndrome (PTPN11 gene mutation positive). J Pediatr Hematol Oncol. 2006;28(3):123–5.

    Article  CAS  PubMed  Google Scholar 

  161. Chatelain EH, Dupuy JW, Letellier T, Dachary-Prigent J. Functional impact of PTP1B-mediated Src regulation on oxidative phosphorylation in rat brain mitochondria. Cell Mol Life Sci. 2011;68(15):2603–13.

    Article  CAS  Google Scholar 

  162. Pagliarini DJ, Wiley SE, Kimple ME, Dixon JR, Kelly P, Worby CA, Casey PJ, Dixon JE. Involvement of a mitochondrial phosphatase in the regulation of ATP production and insulin secretion in pancreatic beta cells. Mol Cell. 2005;19(2):197–207.

    Article  CAS  PubMed  Google Scholar 

  163. Vazquez F, Matsuoka S, Sellers WR, Yanagida T, Ueda M, Devreotes PN. Tumor suppressor PTEN acts through dynamic interaction with the plasma membrane. Proc Nat Acad Sci USA. 2006;103(10):3633–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Zhu Y, Hoell P, Ahlemeyer B, Krieglstein J. PTEN: a crucial mediator of mitochondria-dependent apoptosis. Apoptosis. 2006;11(2):197–207.

    Article  CAS  PubMed  Google Scholar 

  165. Zu L, Zheng X, Wang B, Parajuli N, Steenbergen C, Becker LC, Cai ZP. Ischemic preconditioning attenuates mitochondrial localization of PTEN induced by ischemia-reperfusion. Am J Physiol Heart Circ Physiol. 2011;300(6):H2177–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Liang H, He S, Yang J, Jia X, Wang P, Chen X, Zhang Z, Zou X, McNutt MA, Shen WH, et al. PTENalpha, a PTEN isoform translated through alternative initiation, regulates mitochondrial function and energy metabolism. Cell Metab. 2014;19(5):836–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Endo T, Kohda D. Functions of outer membrane receptors in mitochondrial protein import. Biochim Biophys Acta. 2002;1592(1):3–14.

    Article  CAS  PubMed  Google Scholar 

  168. Meisinger C, Ryan MT, Hill K, Model K, Lim JH, Sickmann A, Muller H, Meyer HE, Wagner R, Pfanner N. Protein import channel of the outer mitochondrial membrane: a highly stable Tom40-Tom22 core structure differentially interacts with preproteins, small tom proteins, and import receptors. Mol Cell Biol. 2001;21(7):2337–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Wiedemann N, Kozjak V, Chacinska A, Schonfisch B, Rospert S, Ryan MT, Pfanner N, Meisinger C. Machinery for protein sorting and assembly in the mitochondrial outer membrane. Nature. 2003;424(6948):565–71.

    Article  CAS  PubMed  Google Scholar 

  170. Rehling P, Model K, Brandner K, Kovermann P, Sickmann A, Meyer HE, Kuhlbrandt W, Wagner R, Truscott KN, Pfanner N. Protein insertion into the mitochondrial inner membrane by a twin-pore translocase. Science. 2003;299(5613):1747–51.

    Article  CAS  PubMed  Google Scholar 

  171. Mesecke N, Terziyska N, Kozany C, Baumann F, Neupert W, Hell K, Herrmann JM. A disulfide relay system in the intermembrane space of mitochondria that mediates protein import. Cell. 2005;121(7):1059–69.

    Article  CAS  PubMed  Google Scholar 

  172. Covian R, Balaban RS. Cardiac mitochondrial matrix and respiratory complex protein phosphorylation. Am J Physiol Heart Circ Physiol. 2012;303(8):H940–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Feng J, Zhu M, Schaub MC, Gehrig P, Roschitzki B, Lucchinetti E, Zaugg M. Phosphoproteome analysis of isoflurane-protected heart mitochondria: phosphorylation of adenine nucleotide translocator-1 on Tyr194 regulates mitochondrial function. Cardiovas Res. 2008;80(1):20–9.

    Article  CAS  Google Scholar 

  174. Grimsrud PA, Carson JJ, Hebert AS, Hubler SL, Niemi NM, Bailey DJ, Jochem A, Stapleton DS, Keller MP, Westphall MS, et al. A quantitative map of the liver mitochondrial phosphoproteome reveals posttranslational control of ketogenesis. Cell Metab. 2012;16(5):672–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Pagliarini DJ, Dixon JE. Mitochondrial modulation: reversible phosphorylation takes center stage? Trend Biochem Sci. 2006;31(1):26–34.

    Article  CAS  PubMed  Google Scholar 

  176. Zhao X, Leon IR, Bak S, Mogensen M, Wrzesinski K, Hojlund K, Jensen ON. Phosphoproteome analysis of functional mitochondria isolated from resting human muscle reveals extensive phosphorylation of inner membrane protein complexes and enzymes. Mol Cell Proteom. 2011;10(1):110000299.

    Article  CAS  Google Scholar 

Download references

Authors’ contributions

SL and TM designed and wrote the manuscript, SL and KS collected and analyzed data; SL, STL, RT, JL and MT revised the manuscript. All authors read and approved the final manuscript.

Acknowledgements

We are Grateful to the support from The Vincent F. Kilborn, Jr. Cancer Research Foundation, National Institutes of Health Grants R01CA149646 (M. Tan), and American heart association scientist development Grant 12SDG10970000 (S.T. Lim).

Competing interests

The authors declare that they have no competing interests.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ming Tan.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lim, S., Smith, K.R., Lim, ST.S. et al. Regulation of mitochondrial functions by protein phosphorylation and dephosphorylation. Cell Biosci 6, 25 (2016). https://doi.org/10.1186/s13578-016-0089-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13578-016-0089-3

Keywords