Zum Hauptinhalt springen

Inhibition of ATR Reverses a Mitochondrial Respiratory Insufficiency

Borror, Megan B. ; Girotti, Milena ; et al.
In: Cells, Jg. 11 (2022-05-01), Heft 11, S. 1731-1731
Online academicJournal

Inhibition of ATR Reverses a Mitochondrial Respiratory Insufficiency 

Diseases that affect the mitochondrial electron transport chain (ETC) often manifest as threshold effect disorders, meaning patients only become symptomatic once a certain level of ETC dysfunction is reached. Cells can invoke mechanisms to circumvent reaching their critical ETC threshold, but it is an ongoing challenge to identify such processes. In the nematode Caenorhabditis elegans, severe reduction of mitochondrial ETC activity shortens life, but mild reduction actually extends it, providing an opportunity to identify threshold circumvention mechanisms. Here, we show that removal of ATL-1, but not ATM-1, worm orthologs of ATR and ATM, respectively, key nuclear DNA damage checkpoint proteins in human cells, unexpectedly lessens the severity of ETC dysfunction. Multiple genetic and biochemical tests show no evidence for increased mutation or DNA breakage in animals exposed to ETC disruption. Reduced ETC function instead alters nucleotide ratios within both the ribo- and deoxyribo-nucleotide pools, and causes stalling of RNA polymerase, which is also known to activate ATR. Unexpectedly, atl-1 mutants confronted with mitochondrial ETC disruption maintain normal levels of oxygen consumption, and have an increased abundance of translating ribosomes. This suggests checkpoint signaling by ATL-1 normally dampens cytoplasmic translation. Taken together, our data suggest a model whereby ETC insufficiency in C. elegans results in nucleotide imbalances leading to the stalling of RNA polymerase, activation of ATL-1, dampening of global translation, and magnification of ETC dysfunction. The loss of ATL-1 effectively reverses the severity of ETC disruption so that animals become phenotypically closer to wild type.

Keywords: aging; ageing; checkpoint response; DNA damage response; DDR; MAK-1; MAK-2; MAPKAPs; Mit mutants; polysome profiling; retrograde response

1. Introduction

The essential nature of mitochondria is underscored by the devastating effects of inherited mitochondrial diseases, which afflict approximately one in 5000 children and almost always result in pathological shortening of life [[1]]. Because these organelles play central roles in many cellular processes, including energy production, nucleotide metabolism, and apoptotic signaling [[2], [4]], their progressive, age-dependent reduction in function means almost all of us, at some time in life, will experience the negative consequences of mitochondrial dysfunction [[5]]. Indeed, a growing number of age-related ailments common to western societies are either associated with, or directly caused by, mitochondrial disruption [[6]].

In Caenorhabditis elegans, as in humans, mitochondrial dysfunction is threshold-dependent, that is, a critical level of electron transport chain (ETC) dysfunction must be reached before pathology is observed [[8]]. Nematodes with high levels of mitochondrial dysfunction exhibit shortened lifespan and reduced fitness, whereas worms with moderate levels of mitochondrial dysfunction display an extended lifespan [[9]]. This response is evolutionarily conserved from yeast to mammals, and understanding the molecular mechanisms regulating lifespan extension in worms in response to mild ETC disruption may provide insight into strategies that could be applied to help keep our own cells above their critical mitochondrial dysfunction threshold.

Reduced ETC activity disrupts many downstream cellular processes, and, as might be expected, cells have evolved a variety of sensors and strategies to counter progression toward their critical ETC threshold. Multiple signaling pathways, collectively called retrograde responses, are activated within cells and these, in turn, control expression of nuclear counter-measures [[10]]. For example, an increase in the unfolded protein load of the mitochondrial matrix re-directs the transcription factor ATFS-1 from the matrix to the nucleus where it stimulates expression of mitochondrial chaperones [[11]]. Other signals, such as changes in calcium concentration, activation of mitogen-activated protein kinases (MAPKs), and even knock-on changes in the cytoplasmic unfolded protein load also result in retrograde response activation [[12], [14]]. Counter-measures that are activated by cells include increased mitochondrial biogenesis, elevated mitochondrial DNA (mtDNA) replication, increased mitophagy, activation of alternate pathways of energy production (such as glycolysis), and changes in the abundance and activity of respiratory complexes or their regulatory factors [[8], [15]]. The complete network of pathways that detect and respond to mitochondrial stress is, however, far from being fully understood [[15]].

Previously, we described a phenomenon in worms in which the loss of CEP-1, a homolog of the human p53 checkpoint protein that recognizes identical DNA sequences and is the only p53 family member present in C. elegans [[16]], right-shifted the mitochondrial ETC threshold, effectively allowing animals to cope with a greater degree of mitochondrial ETC disruption, and consequently mitigating the lifespan-shortening effects of severe mitochondrial dysfunction [[17]]. This response was due to alterations in autophagy and lipid metabolism, both of which are regulated by CEP-1 [[18]]. In the current study, we examined the role of other checkpoint proteins in modulating the mitochondrial threshold effect of worms, since this is still a little-explored area of investigation.

Ataxia-telangiectasia-mutated (ATM) and 'ATM- and rad3-related' (ATR) are two well-studied members of the phosphoinositide 3-kinase-related kinase (PIKK) family of proteins that play essential roles in transducing checkpoint signaling during the DNA damage response (DDR) [[19]]. In humans, both proteins together directly phosphorylate over 900 sites on some 700 proteins, including p53 [[20]]. ATM and ATR differ in the lesions to which they are recruited, with ATM being recruited to dsDNA breaks, and ATR having a broader specificity underscored by the presence of ssDNA, including recessed dsDNA breaks and stalled replication forks [[21]]. Depending on the level of damage and success of repair, checkpoint activation can lead to pro-survival or pro-apoptotic mechanisms [[22]]. Recently, additional functions were ascribed to both ATM and ATR. For ATM, these involve regulation of a variety of processes, including insulin signaling, the pentose phosphate pathway, and mitophagy [[23], [25], [27]]. Moreover, in yeast, ATM/Tel1p was found to act as a specific sensor of mitochondria-generated ROS [[28]]. ATR, on the other hand, was shown to be important in controlling an outer mitochondrial membrane (OMM)-localized apoptotic signal, autophagy, RNA polymerase activity, and chromatin condensation [[28], [30], [32]].

In this study, we uncover a new role for ATR in mitochondrial retrograde response signaling. Studying atl-1(tm853) loss-of-function mutants in C. elegans, we show that ATR/ ATL-1 is activated when the mitochondrial ETC is impaired. We find no evidence for increased nuclear DNA damage under these conditions, yet a significant change in ribose- and deoxyribose nucleotide pools is observed. Consistent with ATL-1 activation being tied to this alteration, we find an elevation in R-loop formation in atl-1(tm853) mutants following exposure to ETC disruption. R-loop formation is indicative of RNA pol II stalling, a known mechanism of ATR activation in human cells. Unexpectedly, we find that mutant atl-1(tm853) worms have half the amount of mitochondrial DNA of wild type worms, yet they show no reduction in the amount of oxygen they consume, neither under control- or ETC-disrupting conditions. In determining the mechanism by which ATL-1 functions to regulate mitochondrial retrograde response signaling, we show that ATL-1 inhibits translation, potentially through ICD-1, a ribosomal exit-tunnel chaperone. Finally, in the process of studying atl-1(tm853) mutants, we have discovered several other DNA damage response proteins, including MAPKAP-2 and MAPKAP-3, that function downstream of mitochondrial ETC disruption. These findings underscore what appears to be an extensive role for the DDR network in mitochondrial retrograde response signaling.

2. Methods

A detailed description of all methods is provided in File S1.

2.1. Nematode Strains and Maintenance

Genotypic information for all strains used in this study is provided in Table S4. Lines were maintained at 20 °C on standard NGM agar plates seeded with E. coli (OP50) [[33]]. To control for possible untoward effects derived from the nT1(IV;V) chromosomal pair in maternally-rescued atl-1(tm853) mutants, all lines, including the wild type control, were moved into the nT1(IV;V) background, and then unbalanced F1 animals were freshly collected for each relevant experiment. Further details are provided in File S1.

2.2. Bacterial Feeding RNAi

All bacterial feeding RNAi constructs were designed in the pL4440 vector backbone and maintained in HT115 bacteria [[34]]. Growth rate, life span, RNAi efficacy assays, and RNAi dilution methods are described in File S1.

2.3. Microscopy

Images of fluorescent worms were captured using an Olympus SZX16 fluorescence microscope connected to an Olympus DP71 CCD camera. In Figure 7C, to determine whether differences existed in the morphology of mitochondria between the four tested conditions, four randomly selected images from each condition were mixed and then grouped by a scorer who was blind to image identity. In a set of 16 images, 10 were co-grouped correctly. The probability that this grouping occurred by chance was modeled using an Excel VBA macro that randomly binned the images into four groups of four, 10,000 times (File S3). The final p-value was <0.0024 and deemed significant.

2.4. mRNA and mtDNA Quantitation

For quantitation of mRNA, fold change in each mRNA of interest was determined using the ΔΔCt method normalized to the geometric mean of cdc-42, pmp-3, and Y45F10D.4 [[35]]. For quantitation of mitochondrial DNA (mtDNA), PCR primer pairs targeting mtDNA-encoded ctb-1 and intron 4 of genomic DNA-encoded ama-1 were used as described [[36]]. The ratio of mitochondrial DNA to nuclear DNA was calculated using Real Time PCR, with ama-1 serving as the nuclear normalization quantity [[37]]. A complete list of qPCR primers is provided in Table S5.

2.5. Mutation Screening Assays

Full methodological details of the unc-93(e1500) and unc-58(e665) reversion assays; the eT1(III;V) and nT1(IV;V) lethal mutation assays; the Lac-Z Frameshift Assay; and the 'terminal deoxynucleotidyl transferase dUTP nick end labeling' (TUNEL) assay are presented in File S1.

2.6. Nucleotide Extraction and Mass Spectrometry

The quantitation of deoxyribonucleotide and ribonucleotide species in whole-worm extracts of isp-1(qm150) and wild-type (N2) animals was undertaken using HPLC-MS at the UTHSCA Mass Spectrometry Facility. Methods for nucleotide extraction, separation, quantitation, and statistical analysis are provided in File S1.

2.7. R-Loop Quantification

R-loop formation in worms was quantified using αS9.6 primary antibody (Kerafast, 1:2000 in 5% Blotto, 16 h, 4 °C), previously shown to be selective for DNA:RNA hybrids [[38]]. A list of R-loop reagents, and detailed methodology describing their use, is provided in File S1.

2.8. Spliceosome Reporter Assays

egl-15 and ret-1 alternate splicing reporter genes were engineered by Kuroyanagi and colleagues [[39]]. Details of reporter construction and their use are provided in File S1.

2.9. Nematode Oxygen Consumption

Nematode oxygen consumption measurements were undertaken using a Seahorse XFe24 Analyzer. Approximately 30 worms were added to each Seahorse well, with 5 replicate wells employed per test condition. We followed the procedure of Luz and colleagues [[41]], and only averaged oxygen consumption rates across the final four measurement phases for each well to provide a single rate value for that well. A total of four independent experiments were performed. Significance testing was undertaken using the Student's t-test, with p < 0.05 considered significant.

2.10. Polysome Profiling

Quantitative polysome profiling was undertaken using the procedure of Steffen and colleagues [[42]] using a Brandel BR-188 Density Gradient Fractionation System. Several modifications were included to optimize for C. elegans polysome extraction, and these are described in detail in File S1. Two fully independent, experimental replicates were collected.

2.11. Western Analysis

Western blotting reagents and suppliers are described in File S1.

2.12. DNA Damage Response (DDR) Screen

The atm-1(gk186); atl-1(tm853) suppressor screen was undertaken using a bacterial feeding RNAi targeting 201 DDR-related genes [[43]]. A protocol describing our screening method, including quantitation of the RNAi effect-size and significance testing, is provided in File S1 and Tables S1–S3.

3. Results

3.1. Loss of ATL-1/ATR Desensitizes Worms to Mitochondrial ETC Stress

In C. elegans, atp-3 encodes the ortholog of the human ATP5O/OSCP subunit of the mitochondrial F1F0 ATP synthase, whereas isp-1 encodes the Rieske Fe-S ortholog of complex III. We have previously reported that wild type worms exposed to increasing amounts of bacterial feeding RNAi targeting either atp-3 or isp-1 reach a critical threshold, after which the concentration of RNAi causes worms to exhibit a reduction in body size and an extension of lifespan [[44]]. For isp-1 RNAi, life extension follows a monotonic function, whereas for atp-3 RNAi, life is first extended, and then it is pathologically shortened [[44]]. We previously mapped these functions using 12 point datasets across multiple rounds of testing [[44]]. The differing effect of atp-3 and isp-1 feeding RNAi on lifespan reflects the more potent degree of ETC inhibition that is ultimately attained following severe atp-3 knockdown.

Using these well-defined treatment regimes, we tested whether loss of the checkpoint proteins ATM-1 and ATR altered the phenotypic response of worms to mitochondrial ETC disruption (see File S1 for a full description of mutants used in this study, and the method used to the propagate lethal DDR mutations). When mutant atm-1(gk186) worms were exposed to atp-3 or isp-1 RNAi (1/10th strength and undiluted), animals matured and survived indistinguishably from similarly-treated wild type worms (Figure 1A–D). In contrast, atl-1(tm853) mutants were differentially refractory to the effects of knockdown of either RNAi. This effect was most evident in worms treated with 1/10th strength atp-3 or isp-1 RNAi, where animals showed obvious resistance to both size reduction (Figure 1A,B) and lifespan extension (Figure 1C,D) (additional examples are presented later in Figure 2A–C). atp-3 and isp-1 knockdown did eventually reduce adult size, and extend life in the mutant atl-1(tm853) background, suggesting the response to ETC disruption in these animals is right-shifted relative to wild type control (see schematic Figure 1C,D). Surprisingly, the measurement of atp-3 and isp-1 mRNA levels following RNAi treatment revealed that target gene knockdown was as efficacious in atl-1(tm853) mutants as it was in wild type worms (Figure 1E), indicating these animals were not deficient for RNAi. Furthermore, we found that atl-1(tm853) mutants were also refractory to mitochondrial ETC disruption induced using chemical inhibition (Figure S1). Based on these findings, we conclude that inhibition of the phosphoinositide 3-kinase-related kinase ATL-1 confers resistance to mitochondrial ETC disruption.

3.2. MAK-1 and MAK-2 Participate in Mitochondrial Retrograde Response Signaling

During the course of our studies, we made the unexpected observation that two independently-generated atm-1(gk186); atl-1(tm853) double mutants regained sensitivity to atp-3 and isp-1 knockdown (Figure 1). This observation provides additional support for the notion that the RNAi machinery is unaffected by the loss of atl-1. To explain this odd finding, we reasoned that ETC disruption might normally induce DNA damage that activates ATR, but not ATM. In the absence of both ATR and ATM, however, secondary forms of DNA damage might accumulate that specifically trigger a third checkpoint response that reinstates sensitivity to atp-3 and isp-1 knockdown. This hypothesis reflects the well-established redundancy in the DNA damage response network [[20]]. Several recent studies by us and others [[13], [36]] have shown a role for MAPK signaling following ETC disruption in worms. Interestingly, the p38 MAPK-activated protein kinases MAPKAP-2 and MAPKAP-3 are checkpoint proteins known in humans to act in parallel to ATM and ATR [[45], [47]]. To explore the possibility that alternate checkpoint signaling is activated in PIKK mutant worms following ETC disruption, we exposed animals to atp-3 knockdown, and simultaneously inhibited production of either MAK-1 or MAK-2, the worm orthologs of MAPKAP-2 and -3, respectively. The effect of these treatments on alt-1(tm853) and atm-1(gk186) single mutants, atm-1(gk186); atl-1(tm853) double mutants, and wild type worms was examined. Our results can be summarized as follows: (i) knockdown of mak-1 or mak-2, either alone or in combination with atp-3 knockdown, had no effect on final adult size in any of the genetic backgrounds (Figure 2A). (ii) Knockdown of mak-1 or mak-2 failed to reproducibly prevent life extension following atp-3 inhibition in the two independent atm-1(gk186); atl-1(tm853) double mutant isolates. (iii) Unexpectedly, the reduced ability of atp-3 RNAi to extend the lifespan of atl-1(tm853) mutants was further impaired when undertaken in conjunction with mak-1 or mak-2. A similar result was also observed for the atm-1(gk186) mutant. (iv) For wild type worms, only mak-1 knockdown significantly impaired atp-3 RNAi-induced life extension (Figure 2B,C). These findings suggest that at least two other checkpoint proteins, namely MAK-1 and MAK-2, are causally involved in lifespan extension (but not size determination) of worms experiencing mitochondrial ETC disruption. The role of different checkpoint proteins is, therefore, dependent upon the functional status of both ATL-1 and ATM-1. When these two checkpoint proteins are disrupted simultaneously, MAK-1 and MAK-2 become dispensable, potentially implying yet another checkpoint system awaits identification in atm-1(gk186); atl-1(tm853) double mutants. Although clearly complicated, this is not unexpected for a response that is network-based and sensitive to a variety of DNA lesions.

3.3. Knockdown of DNA Pol α Reverses the Small Phenotype of atm-1(gk186); atl-1(tm853) Mutant...

To further explore the role of novel checkpoint proteins in modulating the critical ETC threshold in atm-1(gk186); atl-1(tm853) double mutants, we undertook a double RNAi screen of 201 DNA damage response-related genes. We asked whether knockdown of a target gene in this genetic background could re-confer resistance to atp-3 knockdown, similar to atl-1(tm853) mutants (for screen details, see Figure 3A). We identified two genes: scc-3, which encodes a cohesin complex subunit; and Y47D3A.29, which is orthologous to the human PolA1 catalytic subunit of DNA polymerase alpha (DNA pol α, Figure 3B,C). The effect of losing each gene was robust to three different post-hoc tests (each p < 0.001, see Tables S1–S3, File S2, and Methods for details of statistical testing). The isolation of PolA1 is particularly intriguing because it suggests that atp-3 knockdown in atm-1(gk186); atl-1(tm853) double mutants could be inducing some kind of disruption to DNA, its repair, or its replication, and which is bypassable by low-fidelity polymerases (which presumably accommodate knockdown of PolA1). This, in turn, might avoid checkpoint activation and make animals appear more atl-1(tm853)-like in our assay.

3.4. No Evidence for Permanent Nuclear DNA Damage Following Mitochondrial ETC Disruption

Nuclear DNA (nDNA) damage is the canonical signal that results in activation of ATR. In worms, resected ends of dsDNA breaks, and stalled replication forks have both been shown to activate ATL-1 [[48]]. Both lesions contain single-stranded DNA (ssDNA), and binding of heterotrimeric replication protein A (RPA) to these regions, in turn, recruits and activates ATL-1. To test if ETC disruption results in elevated levels of dsDNA breaks, we focused on isp-1(qm150) genetic mutants, since no viable atp-3 mutant exists, and utilized TUNEL staining on sections of paraffin-embedded worms [[49]]. We found no evidence for increased nDNA breakage (Figure S2A). DNA repair is sometimes inaccurate, depending on the repair machinery that is employed. To test if DNA mutation frequency was increased in worms exposed to ETC disruption, we employed four independent genetic assays. To facilitate analysis, these studies were undertaken using atp-3 RNAi. The eT1(III;V) and nT1(IV;V) reciprocal chromosomal translocations each suppress recombination against the regions they balance (~15 Mbp for both balancers). Homozygous-lethal DNA mutations, therefore, accumulate on balanced chromosomes, and they become identifiable when the balancing eT1 chromosomal pair, or the nT1 chromosomal pair, are lost because unbalanced progeny that are homozygous for the mutation are inviable and, hence, absent. Using both assays, we found no evidence for increased nDNA mutation rate following knockdown of atp-3 in wild type worms (Figure 4A). Similarly, the use of two Unc phenotypic reversion assays yielded similar conclusions (Figure 4B). Taken together, our data show there is neither an increase in DNA strand breakage nor is there an increase in nuclear DNA mutation frequency in worms experiencing mitochondrial ETC dysfunction.

3.5. Ribonucleotide and Deoxyribonucleotide Pools Are Disrupted by Mitochondrial ETC Dysfunct...

To search for alternate mechanisms by which ATL-1 might be activated in worms experiencing mitochondrial ETC dysfunction, we examined changes in whole-worm nucleotide levels. As in humans, pyrimidine biosynthesis in C. elegans requires a functional mitochondrial electron transport chain. This is because formation of orotate from dihydroorotate is coupled to reduction of ubiquinone (UQ) in a reaction catalyzed by the inner mitochondrial membrane enzyme, dihydroorotate dehydrogenase (DHODH, Figure S3). In mammals, processes that slow flux through the mitochondrial ETC, downstream of UQ, also slow the rate of dihydroorotate dehydrogenase, resulting in measurable disruption of nucleotide pools [[50]]. Several important enzymes are sensitive to changes in nucleotide abundance, including DNA polymerases which are rate-limited by availability of their deoxyribonucleotide substrates [[51]]. Stalling of DNA polymerase, followed by continued DNA unwinding by the MCM helicase, results in ssDNA formation [[52]], which is a known substrate for ATR activation.

To measure nucleotide levels, we re-focused our attention on the isp-1(qm150) genetic mutant for ease of culture. We interrogated three stages of development—the final two stages of larval development (L3 and L4), and the first day of adulthood. These stages were selected because in wild type worms, there is a heavy demand for nucleotide precursors around the L3/L4 molt, and then again around the L4/adult molt. At these times, the gonad arms are expanding, and there is also a concomitant five-fold and a six-fold increase in mitochondrial DNA (mtDNA), respectively [[54]]. Among nine deoxyribonucleotide species that we were able to resolve and reliably detect (see Methods), we observed significant increases in dCDP and dTMP in L4 larvae of isp-1(qm150) mutants (Figure 5A). The increase in dCDP abundance was maintained into adulthood (Figure 5B). Among eleven ribonucleotide species that we were able to reliably resolve and detect, we observed significant increases in AMP, CMP, and UMP in L4 larvae of isp-1(qm150) mutants (Figure 5C). This was accompanied by a significant decrease in the level of GTP. In adult isp-1(qm150) worms, changes in ribonucleotide pools became more pronounced, with eight species being significantly underrepresented—ADP, ATP, GDP, GTP, CMP, CTP, UDP, and UTP (Figure 5D). Changes in nucleotide pools were specific to L4 and adult worms, since we did not detect any measurable differences in any quantifiable nucleotide species in L3 larvae of isp-1(qm150) mutants relative to wild type worms (Figure 5E). Finally, the absence of a corresponding increase in many of the monophosphate ribonucleotide species in adult isp-1(qm150) worms when their corresponding di- and triphosphate counterparts decreased suggests ribonucleotide pool sizes are reduced in these animals. We combined the levels of each mono-, di-, and triphosphate species, and tested if they differed significantly between strains at each developmental stage, but they did not. Error propagation, however, severely reduced our statistical resolving power (Figure 5F). Collectively, we conclude that disruption of the mitochondrial ETC in C. elegans results in measurable and significant changes in the abundance of multiple nucleotide species, notably with all ribonucleotide triphosphate species becoming underrepresented by adulthood.

3.6. Evidence for RNA Polymerase Stalling Following Mitochondrial ETC Disruption

Stalling of RNA polymerase II (RNA pol II) during transcription is sufficient to activate an ATR, RPA, and p53-dependent DNA damage response in human fibroblasts [[55]]. Given that mitochondrial ETC dysfunction in isp-1(qm150) mutants leads to measurable reductions in all four ribonucleotide triphosphates, and because we have previously shown a role for p53/cep-1 in the life extension of these worms [[17]], we hypothesized that ATL-1 activation following ETC disruption might occur as a result of RNA pol II stalling due to reduced nucleotide availability. We therefore quantified the formation of R-loops on genomic DNA [[56]]. R-loops are DNA:RNA hybrids that accumulate when transcription stalls (Figure 6A). We detected R-loops in purified, whole-worm genomic DNA fractions using slot blot analysis in conjunction with a DNA:RNA hybrid-specific antibody (αS9.6) [[38]]. To obtain sufficient material for these studies, we utilized a double feeding RNAi approach, and tested four conditions: wild type worms exposed to either vector control RNAi, 1/10th strength isp-1 RNAi, 9/10th strength atl-1 RNAi, or both 9/10th strength atl-1 RNAi with 1/10th strength isp-1 RNAi. Vector control feeding RNAi was added as balancer where appropriate. qRT-PCR analysis undertaken before the start of the experiment revealed 9/10th strength atl-1 RNAi reduced atl-1 mRNA by 50% (Figure 6B), the same amount that was observed in maternal-effect rescued atl-1(tm853) mutants (Figure 6C). The results of our R-loop studies can be summarized as follows (Figure 6D): (i) knockdown of isp-1 significantly enhanced R-loop formation relative to vector control (Student's t-test, p < 0.05). (ii) Knockdown of atl-1 did not block R-loop formation by isp-1, as would be expected for a gene that functioned downstream of R-loop accumulation. We next tested whether any differences in DNA:RNA polymer length distinguished R-loops generated in worms exposed to isp-1 RNAi from those exposed to isp-1 and atl-1 double RNAi. Genomic fractions were separated on an agarose gel, stained with ethidium bromide, then transferred to nitrocellulose and probed with αS9.6 antibody (Figure 6E). No overt difference distinguished the two samples; however, in both instances, we noted some of the staining with αS9.6 localized to low molecular weight DNA:RNA fragments (<2 kb), in addition to the higher molecular weight structures that we expected. The identity of these smaller fragments remains uncharacterized (see Section 4). Collectively, our studies suggest that transcriptional stalling is increased in worms exposed to mitochondrial ETC disruption, identifying one route by which ATL-1 is potentially activated in these animals.

3.7. RNA Splicing Is Not Altered by Mitochondrial ETC Disruption

Having identified a possible mechanism by which ATL-1 is activated in worms experiencing mitochondrial ETC stress, we next sought to determine the mode by which ATL-1 acts to further disrupt mitochondrial function. It is well-established that pre-mRNA splicing is a co-transcriptional process [[58]], and it has also been reported that isp-1(qm150) mutants are sensitive to knockdown of two spliceosome factors, SFA-1 and REPO-1, fully ablating their life extension [[59]]. We therefore tested whether the stalling of RNA pol II altered either splicing efficiency or splicing fidelity in worms exposed to ETC disruption, and if ATL-1 played a role in this process. We used two synthetic-gene reporter assays developed by Kuroyanagi and colleagues [[39], [60]], designed to measure alternate splicing at the egl-15 and ret-1 loci, respectively. The latter reporter was shown to be a sensitive marker of age-related splicing dysfunction [[59]]. Each assay has the potential to produce two fluorescent outputs that record changes in the nature of a differential splicing event, and the tissues in which these events occur (Figure S4A,B). We exposed worms carrying the relevant reporter cassette to double feeding RNAi targeting atp-3 or isp-1 (1/10th strength) in conjunction with either atl-1 or vector control (9/10th RNAi strength) (Figure S4C). Animals were followed three days into adulthood. None of the conditions that disrupted the mitochondrial ETC, neither in the presence or absence of atl-1, resulted in alteration of egl-15 or ret-1 splicing relative to vector control-treated animals.

3.8. Reduced ATL-1 Activity Does Not Enhance Hormetic Stress Response Activation

We next tested whether the loss of atl-1 affected the activity of major stress response pathways known to be activated by atp-3 or isp-1 depletion [[3], [13], [44], [61]]. The loss of atl-1 did not result in constitutive activation of Phsp-6:GFP, Pgst-4:GFP, or Ptbb-6:GFP, which are markers of ATFS-1, SKN-1/NRF-2, and PMK-3/p38 activation, respectively (Figure S5A–D); and neither did the loss of atl-1 increase gene expression of sod-3, ugt-61, or hsp-6 (Figure S5E). These results indicate that the reduction of atl-1 does not affect typical mitochondrial stress response pathways.

3.9. Oxygen Consumption Remains Surprisingly Unaltered in atl-1 Mutants Exposed to ETC Disrup...

We next tested whether mitochondria in atl-1(tm853) mutants were inherently different from those of wild type worms. We also analyzed mitochondria from atm-1(gk186) single mutant worms, as well as atl-1(tm853); atm-1(gk186) double mutants. Properties that we examined included mitochondrial DNA (mtDNA) content, nuclear-encoded ETC transcript abundance, mitochondrial morphology, and whole-worm oxygen consumption. Our findings can be summarized as follows: relative to wild type worms, mtDNA copy number is halved in atl-1(tm853) mutants, consistent with an earlier report [[62]]. mtDNA content was also halved in atl-1(tm853); atm-1(gk186) double mutants (Figure 7A). To assess changes in nuclear-encoded ETC transcript abundance, we focused on five genes: nuo-6, mev-1, isp-1, cco-1, and atp-3, representing subunits from mitochondrial complexes I through V, respectively. We found that all transcripts except nuo-6 were significantly decreased relative to wild type worms in all three mutant strains studied. The only exception was mev-1, for which mRNA levels remained unchanged in atm-1(gk186) mutants (Figure 7B).

To quantify changes in mitochondrial morphology, we employed a hexokinase-2:GFP translational reporter (HXK2:GFP) that is localized to the outer mitochondrial membrane [[63]]. We observed a significant increase in the number of fused mitochondria in reporter worms treated with isp-1 RNAi (p < 0.0024, see Methods for details of statistical testing). Co-knockdown of atl-1 did not significantly affect this result. Mitochondria in worms treated only with atl-1 RNAi were indistinguishable from those of wild type animals (Figure 7C). Finally, the most significant finding that we uncovered was related to alterations in oxygen consumption by atl-1(tm853) mutants. When wild type worms, atl-1(tm853) and atm-1(gk186) single mutants, as well as atl-1(tm853); atm-1(gk186) double mutants were cultured on isp-1 RNAi, the basal oxygen consumption rate of atl-1(tm853) mutants, as well as atl-1(tm853); atm-1(gk186) double mutants, significantly increased relative to wild type worms (Figure 7D). Moreover, though isp-1 knockdown uniformly decreased oxygen consumption in wild type worms, atm-1(gk186) single mutants and atl-1(tm853); atm-1(gk186) double mutants, and by the same extent, oxygen consumption in atl-1(tm853) surprisingly remained unchanged following this treatment. We conclude that though atl-1(tm853) mutants contain half as much mtDNA as wild type worms and have a reduced abundance of several nuclear-encoded ETC transcripts, these worms surprisingly maintain normal oxygen consumption, even when confronted with disruption of their mitochondrial electron transport chain.

3.10. Loss of ATR Leads to Recovery of Translational Activity in Worms Exposed to ETC Disrupt...

Several mechanisms could explain how atl-1(tm853) mutants maintain oxygen consumption despite ETC disruption. Any process that ultimately increases flux or efficiency of the respiratory chain could be involved. Enhanced supercomplex formation, elevated matrix Ca[2]+, modification of ETC subunits, or simply enhanced translational output of respiratory subunits could all be behind the unique respiratory response seen in atl-1(tm853) mutants. We focused on translational output because there is evidence from human 293T cells that ATR might directly regulate the ribosomal machinery by affecting phosphorylation of the nascent polypeptide-associated complex (NAC) [[20]]. In worms, ICD-1 is the sole βNAC isoform [[64]]. Although the ATR phosphorylation site of human BTF3/βNAC is not conserved in ICD-1, we observed, by western analysis, that ICD-1 protein levels are selectively elevated in atl-1(tm853) mutants following atp-3 knockdown (isp-1 knockdown was not tested due to limited quantities of antisera, see Section 2) (Figure 8A). Moreover, early analyses involving several hundred microarray datasets [[66]] showed icd-1 formed part of a tightly clustered gene co-expression group (Figure 8B) that is significantly enriched (hypergeometric distribution, p < 0.05) for genes encoding ribosomal and mitochondrial ETC subunits, as well as key translational regulatory factors (Figure 8C). We therefore tested whether global translational activity was downregulated by mitochondrial ETC disruption, and if the loss of atl-1 permitted its recovery. We employed polysome profiling in conjunction with RNAi targeting both isp-1 (1/10th RNAi strength) and either atl-1 or vector (both at 9/10th RNAi strength). We found that the actively translating polysome fraction of worms treated with isp-1 RNAi was decreased by half relative to vector control-treated worms. Strikingly, simultaneous removal of both atl-1 and isp-1 restored polysome activity to wild type levels (Figure 8D). Loss of atl-1 in the absence of ETC disruption did not significantly impact polysome abundance relative to wild type worms. These data imply that ATL-1, when activated by mitochondrial ETC disruption, normally acts to dampen global ribosomal translation. Whether this is through a novel phosphorylation site on ICD-1, the downregulation of factors that normally control icd-1 expression, or via some other means that globally regulates translational activity remains unclear. There must be some specificity in the translational targets of ATR-1, however, because simply elevating global rRNA production by inhibiting ncl-1, a negative regulator of FIB-1/fibrillarin, a nucleolar protein involved in the regulation and maturation of rRNA, is not sufficient to block the size reduction caused by isp-1 bacterial feeding RNAi (Figure 8E,F). Collectively, our results lead to a model for how ATL-1 acts to modulate the mitochondrial threshold effect (Figure 8G).

4. Discussion

In this study, we discovered that the PIKK kinase ATR/ATL-1 plays a specific role in modulating the mitochondrial threshold effect of C. elegans. When flux through the ETC was reduced, ATL-1 was activated. We presented data suggesting that the stalling of RNA polymerase could be the trigger for ATL-1 activation. We recorded measurable reductions in many nucleotide species in adult worms, including substrates for both RNA and DNA polymerase, and we presented biochemical evidence showing RNA polymerase stalling following ETC disruption. In addition to these findings, we showed that ATL-1 paradoxically exacerbates mitochondrial dysfunction during times of mitochondrial stress. Specifically, when worms containing reduced levels of atl-1 were challenged with ETC disruption, unlike wild type animals, these animals showed no signs of impairment in global translational output or respiratory oxygen consumption. Phenotypically, loss of atl-1 effectively reversed a mitochondrial ETC insufficiency in C. elegans. The survival benefits of further dampening ETC activity following an ETC insult are not entirely clear, but dampening of mitochondrial respiratory activity in conjunction with global translational dampening might represent an effort to slow overall metabolism or to maintain stoichiometric balance between nuclear- and mitochondria-encoded ETC proteins [[67]].

4.1. Mode of ATL-1 Activation Following Mitochondrial ETC Disruption

In our search for signals that triggered ATL-1 activation in response to ETC disruption, we found no evidence for elevated rates of nuclear DNA mutation or strand breakage. This does not preclude the possibility that DNA damage occurred and was accurately repaired. Such damage would have escaped detection in our assays. Nonetheless, we investigated other potential mechanisms of ATL-1 activation, and, in isp-1(qm150) mutants, found that nucleotide pools were significantly changed. This is significant because ATR activation has been linked to changes in nucleotide pools in other model systems. Specifically, in S. cerevisiae, glucose starvation induces Snf1p (AMP kinase) at the mitochondrial surface following reduction of adenosine ribonucleotide energy charge. In turn, Snf1p was shown to phosphorylate Mec1p (ATR) [[32]]. Whether a similar mode of ATR activation functions in C. elegans remains unclear, but Ggc1p, the protein that initially recruits Mec1p to the yeast mitochondrial membrane under low-glucose conditions, has no worm ortholog. We pursued an alternate hypothesis for nucleotide-dependent activation of ATR in worms, namely the stalling of RNA and/or DNA polymerases.

Previously, it had been shown that blocking RNA polymerase during transcriptional elongation is sufficient to activate ATR and induce a DNA damage response (DDR) [[55]]. Indeed, we found evidence that DNA:RNA R-loop hybrids are increased in the presence of mitochondrial ETC disruption, which is indicative of transcriptional stalling. However, it is unclear whether all the R-loops we detected were genomic DNA (gDNA)-derived transcription intermediates. When examining samples on an agarose gel in conjunction with an antibody targeting these structures (S9.6), R-loops are expected to appear at a high molecular weight in association with the gDNA band. We clearly detected an elevated signal in this region in our isp-1 RNAi-treated samples, consistent with our hypothesis. Moreover, we observed an additional smear of low molecular weight nucleic acids. At this point, it is unknown if these low molecular weight molecules are truly DNA:RNA hybrids or a contaminating nucleic acid. The S9.6 antibody has a five-fold higher affinity for DNA:RNA hybrids over RNA:RNA hybrids [[38]], so in anticipation of the latter, our samples were treated with RNase A, ruling out this possibility. One idea is that the low molecular weight DNA:RNA hybrids are derived from mtDNA fragments that have leaked from dysfunctional mitochondria. Consistent with this idea, in mouse embryonic fibroblasts, mitochondrial stress results in mtDNA leaking into the cytoplasm and initiating an immune response [[68]]. On the other hand, worms use a rolling circle replication method for their mtDNA, so precisely how R-loops would form in the absence of D-loop (theta) replication, as is used by mammalian cells, remains enigmatic [[69]]. Another idea is that the low molecular weight species we detected are cytoplasmic DNA:RNA hybrids produced by DNA polymerase α (DNA pol α). Starokadomsky and colleagues [[70]] showed that a small fraction of DNA pol α localizes to the cytoplasm in human cells, where it generates DNA:RNA hybrids. These molecules directly modulate production of type I interferons, and possibly exist to control the activation threshold of the innate immune response by foreign nucleic acids. In patients with X-linked reticulate pigmentary disorder (XLPDR), a primary immunodeficiency with autoinflammatory features, levels of the catalytic subunit of DNA pol α (DNA PolA1) are decreased enough to disrupt cytoplasmic DNA:RNA hybrid formation, but not enough to affect DNA replication [[70]]. Intriguingly, in our screen for DDR proteins that, when inhibited, reversed the small size phenotype of atl-1(tm853); atm-1(gk186) double mutants experiencing mitochondrial ETC stress, we identified an RNAi clone targeting DNA PolA1. Although not discussed earlier, we also isolated a second DNA pol α clone in that screen that targeted a different region of DNA PolA1, but instead resulted in synthetic lethality. This dose-dependent phenotypic response is remarkably similar to what was just discussed for humans. Obviously, more work is needed to determine the identity, provenance, and significance of the low molecular weight DNA:RNA species in worms experiencing ETC stress, and if and how they might relate to ATR activation.

4.2. Downstream Targets of ATL-1 That Modulate Mitochondrial ETC Activity

One important finding from our study is the observation that atl-1 knockdown mitigates the effects of mitochondrial ETC dysfunction. This is an intriguing finding, as it suggests the normal cellular response to ETC disruption is ostensibly further detrimental. Two possible explanations for how the loss of atl-1 could result in enhanced ETC function are: atl-1 knockdown triggers active restoration of mitochondrial function, or, alternatively, atl-1 knockdown precludes propagation of whatever negative downstream effects are normally elicited by ETC disruption. According to our data, atl-1(tm853) mutants contain half the amount of mtDNA as wild type worms, and have a significantly reduced abundance of nuclear-encoded ETC transcripts. These findings argue against the first hypothesis. Surprisingly, despite their shortcomings, we find that atl-1(tm853) mutants are still able to maintain oxygen consumption at near normal levels when confronted with ETC disruption. The precise mechanism involved is yet to be determined, but it may be as simple as atl-1(tm853) mutants being able to maintain the translation of a factor that enhances supercomplex formation, for example [[71]]. Alternatively, atl-1(tm853) mutants may be able to keep their cells fortified with sufficient ETC proteins to counteract any rapid turnover that is likely to occur when subunits become rate-limiting during the assembly process. Along these lines, when we examined polysome formation under conditions of ETC dysfunction, we found that animals with functional ATL-1 had reduced polysome formation, but removal of atl-1 restored this number back to pre-ETC disruption levels. Future studies will be aimed at determining the precise mRNA species present in these polysomes.

Another way in which atl-1(tm853) mutants might be able to counter ETC disruption involves the nascent protein-folding machinery. We showed that ICD-1/βNAC, a protein that acts as a ribosome-associated chaperone, is upregulated in atl-1 mutants upon ETC disruption. In yeast and plants, βNAC plays a role in targeting proteins to the mitochondria, and is required for efficient mitochondrial function [[72], [74], [76]]. βNAC was also one of the few proteins found to be specifically regulated by ATR-dependent phosphorylation in mammals [[20]]. The ATR-dependent phosphorylation site in mammals is not conserved in C. elegans; however, there are several other potential phosphorylation sites in the ICD-1 protein sequence [[77]]. It will be of particular interest in the future to ascertain if ATR directly regulates βNAC abundance on ribosomes to modulate translation of ETC transcripts in worms.

4.3. Mitochondrial Dysfunction in Mammals

Johnson and colleagues [[78]] successfully extended the survival of a mouse model of Leigh syndrome, which was deficient in Ndufs4 of complex I, by chronic administration of the mTOR inhibitor, rapamycin. mTOR is a positive regulator of p70S6 kinase, and a negative regulator of 4E-BP, both of which work to elevate translation. At first, this seems surprising in light of our findings, suggesting global translation needs to be elevated to counter ETC dysfunction in worms. However, though acute inhibition of mTOR indeed results in a global dampening of translation, it also results in the specific translational upregulation of mitochondrial ETC components [[79]]. Moreover, it has been reported that in contrast to acute rapamycin administration, chronic rapamycin administration does not lead to reduced ribosomal activity in vivo [[81]]. Whether the increase in polysomes that we observed in atl-1 animals exposed to ETC disruption represented upregulation of global mRNA translation or just a discrete subset of mRNAs only encoding ETC proteins remains to be determined.

In conclusion, we have shown that the blockade of atl-1 effectively reverses ETC disruption in C. elegans. Preventing mitochondria from transitioning into their critical ETC threshold appears, quite literally, to be the difference between life and death.

Figures

Graph: Figure 1 Loss of ATL-1 desensitizes worms to mitochondrial respiratory chain stress. (A,B) atl-1(tm853) knockout mutants are resistant to size reduction induced by atp-3 or isp-1 bacterial feeding RNAi (panel A). Tested worms include wild type, atm-1(gk186) and atl-1(tm853) single mutants, as well as two independent atl-1(tm853); atm-1(gk186) lines. Worm length is quantified in panel B. All worms were seeded at the same time. Each row was photographed at a common time interval after seeding. This interval increased with increasing RNAi potency, since worms matured more slowly. Evident across rows are differences in growth rates when strains are cultured on the same RNAi treatment. In (B), error bars refer to SEM. Asterisks refer to significant difference relative to respective wild type sample (Students' t-test): ** p < 0.01, *** p < 0.001. (C,D) Charts on left: Adult lifespan of worm strains tested in (A). Shown is a representative study. Samples listed in each table were tested in parallel. For each strain, a synchronous population of 8000 eggs was distributed over the three relevant test conditions. After hatching, unbalanced worms were retrieved (~1/16th). Replicate data for atp-3 are presented in Figure 2. Right panels: (Top) Box plots showing lifespan distribution of selected worm populations from chart on left. Box shows mean +/−25%. Bars: 5–95%. (Bottom) Schematics illustrating the right-shifting effect of atl-1 removal on mean lifespan following exposure to atp-3 and isp-1 feeding RNAi. Curve shapes are based on the 12-point RNAi data collected in Rea et al. [[44]]. (E) RNAi knockdown efficacy is not altered in atl-1(tm853) mutants. mRNA levels were quantified by q-RT PCR following exposure to 1/10th strength atp-3 or isp-1 bacterial feeding RNAi (n = 3 and n = 1 experimental replicates, respectively). Error bars: SEM. Significance testing: Student's t-test. Asterisks refer to difference relative to wild type vector control: ** p < 0.01, *** p < 0.001. In (A–E), the wild type control line (SLR4) was deliberately moved into the nT1 genetic background as the appropriate control (see Section 2).

Graph: Figure 2 MAPKAP kinases are active following mitochondrial respiratory chain stress. (A–C) The MAPKAP kinases MAK-1 and MAK-2 play no role in size control of worms exposed to 1/10th strength atp-3 feeding RNAi (A); however, they do control lifespan in wild type worms and in atm-1 (gk186) and atl-1(tm853) single gene mutants (B,C). Worms in (A,B) were established independently using a total of 8000 and 32,000 synchronized eggs, respectively, and all conditions in each panel were tested in parallel. In (A), worms were photographed at the same chronological point. In (B), the vertical line marks median survival of wild type worms on vector control RNAi. In (A–C), all worms are the unbalanced F1 progeny of parents carrying the nT1 reciprocal chromosomal translocation. * Full genotypic information provided in Table S4.

Graph: Figure 3 Screen for additional checkpoint proteins activated by mitochondrial ETC dysfunction. (A) Schematic of RNAi screen used to identify DNA damage response genes active in atl-1(tm853); atm-1(gk186) double mutants following atp-3 knockdown. Test RNAi that differentially countered the size-reducing effect of 1/10th strength atp-3 RNAi in unbalanced atl-1(tm853); atm-1(gk186) worms, but not nT1-containing atl-1(tm853); atm-1(gk186)/+ worms, were sought (hit). (B) Results of RNAi screen. The effect of each test RNAi on the length of unbalanced worms and nT1-containing worms in the presence and absence of 1/10th strength atp-3 RNAi is reported as a mean difference ratio. The 95% confidence interval (CI) for each ratio is shown. RNAi knockdown of scc-3 and Y47D3A.29 (labeled DNA PolA1) both significantly reduced the mean ratio relative to vector-only test RNAi (p < 0.001, see Methods for significance testing procedure). (C) Top row: Size distribution of worm populations treated with scc-3(left) or Y47D3A.29 (right) feeding RNAi. Bottom row: Histograms of mean difference ratios for worm populations treated with scc-3(left) or Y47D3A.29 (right) feeding RNAi, and plotted relative to worms treated with vector in place of test RNAi. Mean difference ratios were generated by random sampling from the distributions in the top row.

Graph: Figure 4 Nuclear DNA mutation rate is not increased following mitochondrial ETC disruption. (A,B) Multiple genetic screens reveal no significant increase in nuclear DNA mutation frequency following knockdown of atp-3 (p > 0.05, Fisher's exact test). Phenotypes scored include lethal nuclear DNA mutation events covered by the eT1 and nT1 chromosomal translocations (A), and recovery of wild type movement in unc-58(3665) and unc-93(e1500) mutants (B). Knockdown of replication protein A (rpa-1) was included as a positive control in (B).

Graph: Figure 5 Ribonucleotide and deoxyribonucleotide pool ratios are disrupted following mitochondrial ETC disruption. (A–E) Absolute levels of deoxyribonucleotide (A,C) and ribonucleotide (B,D) species in wild type (N2) and isp-1(qm150) worms at the L4 and day-one adult stages, measured using quantitative LC-MS. Red asterisks indicate nucleotide quantities that differ significantly between strains (Student's t-test, 5% FDR). Four nucleotides were not measurable by our assay (dGMP, dGDP, dGTP, GMP). No significant differences in any nucleotide level existed between strains at the L3 larval stage (E). (F) Relative nucleotide pool sizes in L3, L4 and day-one adult N2 and isp-1(qm150) worms, determined by summing together the mono-, di-, and triphosphate species of each deoxyribonucleotide or ribonucleotide (guanosine was an exception) and normalizing to dCNP or CNP, respectively, across each larval stage. No significance difference in each respective nucleotide pool size between strains was detected (Student's t-test, 5% FDR; n = 3, 7 and 3 experimental replicates for N2 L3, L4, and adult worms; and n = 3, 5, and 3 for isp-1(qm150) L3, L4, and adult worms, respectively).

Graph: Figure 6 Evidence for RNA Pol II stalling following mitochondrial ETC stress. (A) Schematic showing genesis of R-loops following stalling of RNA Polymerase II. Modified from [[57]]. (B,C) RNAi knockdown of atl-1 in wild type (N2) worms reduces atl-1 transcript abundance to the same extent as in maternal-effect atl-1(tm853) mutants. In (C), all worms are unbalanced F1 progeny from parents carrying the nT1 reciprocal chromosomal translocation (n = 3; error bars: SEM; significance testing: Student's t-test; symbols: significantly different from wild type: * p < 0.05; *** p < 0.001; significantly different from atl-1(tm853): ^ p < 0.05). In (B), n = 2 and # marks range. (D) R-loops are significantly increased in N2 worms following knockdown of isp-1 (1/10th RNAi strength). Combined knockdown of isp-1 (1/10th RNAi strength) and atl-1 (9/10th RNAi strength) does not abrogate R-loop accumulation. Top panel: representative immunoblot of R-loops; bottom panel: immunoblot quantitation: n = 3; error bars: SEM; Student's t-test (vs. vector): * p ≤ 0.05, ** p ≤ 0.01. (E) Samples in (D) were further analyzed by agarose gel electrophoresis (top two panels), and then western blotting using an antibody that cross reacts with DNA:RNA hybrids (αS9.6, bottom panel). Knockdown of isp-1 results in an increase of gDNA-associated R-loops and unknown small molecular weight DNA:RNA hybrids.

Graph: Figure 7 atl-1(tm853) mutants maintain normal oxygen consumption when confronted with disruption of their mitochondrial ETC. (A) Mitochondrial DNA abundance is significantly decreased in atl-1(tm853) mutants compared to wild type worms (n = 3). (B) Transcript abundance of nuclear-encoded ETC genes is significantly decreased in atl-1(tm853) mutants relative to wild type worms (n = 3; Cx: complex). (C) Morphology of body-wall muscle mitochondria assessed using HXK-2:GFP [[63]], following atl-1 knockdown in the absence or presence of mitochondrial ETC disruption (1/10th strength isp-1 RNAi). Both atl-1 and isp-1 knockdown significantly alter mitochondrial morphology relative to vector-control (p < 0.0024, refer to Methods for significance testing). Scale bar: 20 microns. (D) atl-1(tm853) mutants are resistant to changes in oxygen consumption following isp-1 knockdown (1/10th strength RNAi). All other tested strains experienced a significant decrease in oxygen consumption (N = 4). Oxygen consumption was measured using a Seahorse XFe24 Analyzer. In (A,D), all worms are the unbalanced F1 progeny of parents carrying the nT1 reciprocal chromosomal translocation. In (A–D), asterisks indicate a significant difference from wild type (Student's t-test) unless otherwise noted by a bar for the relevant comparison group: ** p < 0.01. *** p < 0.001; ns, not significant. In (D), ^^^ p < 0.001 in comparison to wild type on isp-1 RNAi. In (A,B), error bars are SEM; in (D), box and whisker plots show 10–90 percentile, with outliers marked (dots).

MAP: Figure 8 Knockdown of atl-1 prevents global translation reduction caused by mitochondrial ETC disruption. (A) Top panel: ICD-1 protein abundance in atl-1(tm853) mutants relative to wild type animals following atp-3 knockdown (N = 2, bars indicate range). Bottom panel: Representative western blot. All worms are the unbalanced F1 progeny of parents carrying the nT1 reciprocal chromosomal translocation. (B) C. elegans gene co-expression terrain map generated using VxInsight [[66]] showing icd-1-containing 'Mount 23' (white arrow). Progressive zoom shots of Mount 23 are shown (left to right). The position of icd-1 is marked by a white square. (C) 88 of 133 (66%) genes comprising Mount 23 of the C. elegans gene co-expression terrain map encode either mitochondrial ETC proteins or ribosomal proteins, including both α- and β-NAC (ICD-1). (D) Polysome profiling shows knockdown of atl-1 (9/10th RNAi strength) restores the level of actively-translating ribosomes to control levels in worms grown on isp-1 RNAi (1/10th strength). (E) ncl-1(e1865) and ncl-1(e1942) mutants do not confer resistance to the size-reducing effects of bacterial feeding RNAi targeting isp-1 relative to wild type (N2) worms. ncl-1 mutants are naturally larger that N2 worms. (mean +/− SEM). (F) RNAi remains efficacious in ncl-1(e1865) and ncl-1(e1942) mutants. Shown is survival following exposure to increasing doses of a lethal feeding RNAi (skn-1). Data from a single experiment are shown that were established simultaneously with common reagents and using 20 synchronous one-day-old adult worms per test condition. (G) Model showing how removal of ATL-1 counters loss of mitochondrial electron transport chain disruption. ETC insufficiency results in nucleotide imbalances, stalling of nucleotide polymerases, activation of ATL-1, and dampening of global translation, leading to magnification of ETC dysfunction.

Author Contributions

Conceptualization, M.B.B., N.V. and S.L.R.; methodology, M.K.C., V.L.M. and X.G.; software, S.L.R.; formal analysis, M.G., M.K.C. and S.L.R.; investigation, M.B.B., A.K., B.H., S.B. (Shylesh Bhaskaran), S.B. (Sandra Becerra), N.N., N.V. and S.L.R.; writing—original draft preparation, M.B.B. and S.L.R.; writing—review and editing, all authors.; supervision, T.E.J., B.K.K., S.L.R.; project administration, S.L.R.; funding acquisition, M.B.B., T.E.J., B.K.K., S.L.R. All authors have read and agreed to the published version of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data presented in this study are contained within the article or Supplemental Materials. Raw data files are available upon request from the corresponding author (SLR).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ATMataxia telangiectasia mutated
ATFS-1activating transcription factor associated with stress
ATL-1ataxia telangiectasia mutated-like (C. elegans ortholog of ATR)
ATRataxia telangiectasia mutated and Rad3-related
DDRDNA damage response
DHODHdihydroorotate dehydrogenase
DNA pol αDNA polymerase alpha
DNA polA1DNA polymerase alpha catalytic subunit
ETCelectron transport chain
mtDNAmitochondrial DNA
MAPKAPp38 MAPK-activated protein kinase
mTORmechanistic target of rapamycin
PIKKphosphoinositide 3-kinase-related kinase
R-loopDNA:RNA hybrid
RNA pol IIRNA polymerase II
TUNELterminal deoxynucleotidyl transferase dUTP nick end labeling

Acknowledgments

We thank the following people: Hidehito Kuroyanagi (Tokyo Medical and Dental University) for the splice reporters; Joel Rothman (UCSB) for the icd-1 antibody; Alison Kell (CU, Boulder) for assistance with DDR library construction; Oxana Radetskaya (UW) for critical reading of the manuscript. Technical assistance was provided by Rebecca Lane, Carlos Galeano, Yvonne Penrod, Haley Beam, Bill Swan, and Melissa Ligon (all UT Health San Antonio). We also thank Wormbase. Some strains were provided by the Caenorhabditis Genetic Stock Center, which is funded by NIH Office of Research Infrastructure Programs (P40 OD010440).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells11111731/s1, Figure S1. Loss of ATL-1 desensitizes worms to mitochondrial respiratory chain stress induced by ethidium bromide. Figure S2. DNA mutation detection in C. elegans using β-Gal frame-shift and TUNEL assays. Figure S3. Nucleotide biosynthesis in C. elegans highlighting the role of the mitochondrial ETC. Figure S4. mRNA splicing remains unaltered in worms experiencing ETC disruption, both in the absence or presence of atl-1 knockdown. Figure S5. atl-1 knockdown does not alter retrograde response pathways normally induced by mitochondrial ETC stress. Table S1. Summary of descriptive statistics for final-round DDR test RNAi screen hits (related to Figure 3). Table S2. Mean ranks of final-round DDR RNAi hits used for Kruskal–Wallis test (related to Figure 3). Table S3. Post-hoc tests for each final-round DDR RNAi hit versus vector (related to Figure 3). Table S4. List of C. elegans strains employed in current study. Table S5. qPCR primer list. File S1. Supplemental Methods and Supplemental Figure Legends. File S2. Mean difference ratios of all hits from DNA Damage Response (DDR) screen (related to Figure 3). File S3. Significance testing for differences in HXK2:GFP fluorescence (related to Figure 7C). References [[82], [84], [86], [88], [90], [92]] are cited in Supplementary Materials.

Footnotes 1 Publisher's Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. References Schaefer A.M., Taylor R.W., Turnbull D.M., Chinnery P.F. The epidemiology of mitochondrial disorders—Past, present and future. Biochim. Biophys. Acta (BBA)-Bioenerg. 2004; 1659: 115-120. 10.1016/j.bbabio.2004.09.005. 15576042 2 Li M.X., Dewson G. Mitochondria and apoptosis: Emerging concepts. F1000Prime Rep. 2015; 7: 42. 10.12703/P7-42. 26097715 3 Liu Y., Samuel B.S., Breen P.C., Ruvkun G. Caenorhabditis elegans pathways that surveil and defend mitochondria. Nature. 2014; 508: 406-410. 10.1038/nature13204 4 Khutornenko A.A., Roudko V.V., Chernyak B.V., Vartapetian A.B., Chumakov P.M., Evstafieva A.G. Pyrimidine biosynthesis links mitochondrial respiration to the p53 pathway. Proc. Natl. Acad. Sci. USA. 2010; 107: 12828-12833. 10.1073/pnas.0910885107 5 Wallace D.C. Mitochondrial DNA mutations in disease and aging. Environ. Mol. Mutagen. 2010; 51: 440-450. 10.1002/em.20586. 20544884 6 Wallace D.C. Mitochondria and cancer. Nat. Rev. Cancer. 2012; 12: 685-698. 10.1038/nrc3365 7 Lane R.K., Hilsabeck T., Rea S.L. The role of mitochondrial dysfunction in age-related diseases. Biochim. Biophys. Acta (BBA)-Bioenerg. 2015; 1847: 1387-1400. 10.1016/j.bbabio.2015.05.021. 26050974 8 Rossignol R., Faustin B., Rocher C., Malgat M., Mazat J.P., Letellier T. Mitochondrial threshold effects. Biochem. J. 2003; 370: 751-762. 10.1042/bj20021594 9 Munkácsy E., Rea S.L. The paradox of mitochondrial dysfunction and extended longevity. Exp. Gerontol. 2014; 56: 221-233. 10.1016/j.exger.2014.03.016 Arnould T., Michel S., Renard P. Mitochondria Retrograde Signaling and the UPR(mt): Where Are We in Mammals?. Int. J. Mol. Sci. 2015; 16: 18224-18251. 10.3390/ijms160818224 Nargund A.M., Pellegrino M.W., Fiorese C.J., Baker B.M., Haynes C.M. Mitochondrial import efficiency of ATFS-1 regulates mitochondrial UPR activation. Science. 2012; 337: 587-590. 10.1126/science.1223560. 22700657 Biswas G., Adebanjo O.A., Freedman B.D., Anandatheerthavarada H.K., Vijayasarathy C., Zaidi M., Kotlikoff M., Avadhani N.G. Retrograde Ca2+ signaling in C2C12 skeletal myocytes in response to mitochondrial genetic and metabolic stress: A novel mode of inter-organelle crosstalk. EMBO J. 1999; 18: 522-533. 10.1093/emboj/18.3.522. 9927412 Munkacsy E., Khan M.H., Lane R.K., Borror M.B., Park J.H., Bokov A.F., Fisher A.L., Link C.D., Rea S.L. DLK-1, SEK-3 and PMK-3 Are Required for the Life Extension Induced by Mitochondrial Bioenergetic Disruption in C. elegans. PLoS Genet. 2016; 12e1006133. 10.1371/journal.pgen.1006133. 27420916 Heo J.M., Livnat-Levanon N., Taylor E.B., Jones K.T., Dephoure N., Ring J., Xie J., Brodsky J.L., Madeo F., Gygi S.P. A stress-responsive system for mitochondrial protein degradation. Mol. Cell. 2010; 40: 465-480. 10.1016/j.molcel.2010.10.021. 21070972 Jiang P., Jin X., Peng Y., Wang M., Liu H., Liu X., Zhang Z., Ji Y., Zhang J., Liang M. The exome sequencing identified the mutation in YARS2 encoding the mitochondrial tyrosyl-tRNA synthetase as a nuclear modifier for the phenotypic manifestation of Leber's hereditary optic neuropathy-associated mitochondrial DNA mutation. Hum. Mol. Genet. 2016; 25: 584-596. 10.1093/hmg/ddv498. 26647310 Huyen Y., Jeffrey P.D., Derry W.B., Rothman J.H., Pavletich N.P., Stavridi E.S., Halazonetis T.D. Structural Differences in the DNA Binding Domains of Human p53 and Its C. elegans Ortholog Cep-1. Structure. 2004; 12: 1237-1243. 10.1016/j.str.2004.05.007 Ventura N., Rea S.L., Schiavi A., Torgovnick A., Testi R., Johnson T.E. p53/CEP-1 increases or decreases lifespan, depending on level of mitochondrial bioenergetic stress. Aging Cell. 2009; 8: 380-393. 10.1111/j.1474-9726.2009.00482.x Schiavi A., Torgovnick A., Kell A., Megalou E., Castelein N., Guccini I., Marzocchella L., Gelino S., Hansen M., Malisan F. Autophagy induction extends lifespan and reduces lipid content in response to frataxin silencing in C. elegans. Exp. Gerontol. 2013; 48: 191-201. 10.1016/j.exger.2012.12.002 Lovejoy C.A., Cortez D. Common mechanisms of PIKK regulation. DNA Repair. 2009; 8: 1004-1008. 10.1016/j.dnarep.2009.04.006 Matsuoka S., Ballif B.A., Smogorzewska A., McDonald E.R. III, Hurov K.E., Luo J., Bakalarski C.E., Zhao Z., Solimini N., Lerenthal Y. ATM and ATR Substrate Analysis Reveals Extensive Protein Networks Responsive to DNA Damage. Science. 2007; 316: 1160-1166. 10.1126/science.1140321 Marechal A., Zou L. DNA damage sensing by the ATM and ATR kinases. Cold Spring Harb. Perspect. Biol. 2013; 5: a012716. 10.1101/cshperspect.a012716. 24003211 Vousden K.H., Lane D.P. p53 in health and disease. Nat. Rev. Mol. Cell Biol. 2007; 8: 275-283. 10.1038/nrm2147. 17380161 Yang D.Q., Kastan M.B. Participation of ATM in insulin signalling through phosphorylation of eIF-4E-binding protein 1. Nat. Cell Biol. 2000; 2: 893-898. 10.1038/35046542. 11146653 Schneider J.G., Finck B.N., Ren J., Standley K.N., Takagi M., Maclean K.H., Bernal-Mizrachi C., Muslin A.J., Kastan M.B., Semenkovich C.F. ATM-dependent suppression of stress signaling reduces vascular disease in metabolic syndrome. Cell Metab. 2006; 4: 377-389. 10.1016/j.cmet.2006.10.002 Cosentino C., Grieco D., Costanzo V. ATM activates the pentose phosphate pathway promoting anti-oxidant defence and DNA repair. EMBO J. 2011; 30: 546-555. 10.1038/emboj.2010.330 Schroeder E.A., Raimundo N., Shadel G.S. Epigenetic silencing mediates mitochondria stress-induced longevity. Cell Metab. 2013; 17: 954-964. 10.1016/j.cmet.2013.04.003 Fang E.F., Bohr V.A. NAD+: The convergence of DNA repair and mitophagy. Autophagy. 2017; 13: 442-443. 10.1080/15548627.2016.1257467 Chiolo I., Minoda A., Colmenares S.U., Polyzos A., Costes S.V., Karpen G.H. Double-strand breaks in heterochromatin move outside of a dynamic HP1a domain to complete recombinational repair. Cell. 2011; 144: 732-744. 10.1016/j.cell.2011.02.012 Gupta A., Sharma S., Reichenbach P., Marjavaara L., Nilsson A.K., Lingner J., Chabes A., Rothstein R., Chang M. Telomere length homeostasis responds to changes in intracellular dNTP pools. Genetics. 2013; 193: 1095-1105. 10.1534/genetics.112.149120 Kumar A., Mazzanti M., Mistrik M., Kosar M., Beznoussenko G.V., Mironov A.A., Garre M., Parazzoli D., Shivashankar G.V., Scita G. ATR mediates a checkpoint at the nuclear envelope in response to mechanical stress. Cell. 2014; 158: 633-646. 10.1016/j.cell.2014.05.046 Hilton B.A., Li Z., Musich P.R., Wang H., Cartwright B.M., Serrano M., Zhou X.Z., Lu K.P., Zou Y. ATR Plays a Direct Antiapoptotic Role at Mitochondria, which Is Regulated by Prolyl Isomerase Pin1. Mol. Cell. 2015; 60: 35-46. 10.1016/j.molcel.2015.08.008. 26387736 Yi C., Tong J., Lu P., Wang Y., Zhang J., Sun C., Yuan K., Xue R., Zou B., Li N. Formation of a Snf1-Mec1-Atg1 Module on Mitochondria Governs Energy Deprivation-Induced Autophagy by Regulating Mitochondrial Respiration. Dev. Cell. 2017; 41: 59-71.e4. 10.1016/j.devcel.2017.03.007. 28399401 Wood W.B.. The Nematode Caenorhabditis elegans; Cold Spring Harbor Laboratory: New York, NY, USA. 1988: 667 Timmons L., Fire A. Specific interference by ingested dsRNA. Nature. 1998; 395: 854. 10.1038/27579. 9804418 Hoogewijs D., Houthoofd K., Matthijssens F., Vandesompele J., Vanfleteren J.R. Selection and validation of a set of reliable reference genes for quantitative sod gene expression analysis in C. elegans. BMC Mol. Biol. 2008; 99. 10.1186/1471-2199-9-9 Sugimoto T., Mori C., Takanami T., Sasagawa Y., Saito R., Ichiishi E., Higashitani A. Caenorhabditis elegans par2.1/mtssb-1 is essential for mitochondrial DNA replication and its defect causes comprehensive transcriptional alterations including a hypoxia response. Exp. Cell Res. 2008; 314: 103-114. 10.1016/j.yexcr.2007.08.015 Rooney J.P., Ryde I.T., Sanders L.H., Howlett E.H., Colton M.D., Germ K.E., Mayer G.D., Greenamyre J.T., Meyer J.N. PCR based determination of mitochondrial DNA copy number in multiple species. Methods Mol. Biol. 2015; 1241: 23-38. 10.1007/978-1-4939-1875-1_3 Zhang Z.Z., Pannunzio N.R., Hsieh C.L., Yu K., Lieber M.R. Complexities due to single-stranded RNA during antibody detection of genomic RNA:DNA hybrids. BMC Res. Notes. 2015; 8127. 10.1186/s13104-015-1092-1 Kuroyanagi H., Watanabe Y., Suzuki Y., Hagiwara M. Position-dependent and neuron-specific splicing regulation by the CELF family RNA-binding protein UNC-75 in Caenorhabditis elegans. Nucleic Acids Res. 2013; 41: 4015-4025. 10.1093/nar/gkt097 Kuroyanagi H., Ohno G., Sakane H., Maruoka H., Hagiwara M. Visualization and genetic analysis of alternative splicing regulation in vivo using fluorescence reporters in transgenic Caenorhabditis elegans. Nat. Protoc. 2010; 5: 1495-1517. 10.1038/nprot.2010.107 Luz A.L., Rooney J.P., Kubik L.L., Gonzalez C.P., Song D.H., Meyer J.N. Mitochondrial Morphology and Fundamental Parameters of the Mitochondrial Respiratory Chain Are Altered in Caenorhabditis elegans Strains Deficient in Mitochondrial Dynamics and Homeostasis Processes. PLoS ONE. 2015; 10e0130940. 10.1371/journal.pone.0130940 Steffen K.K., MacKay V.L., Kerr E.O., Tsuchiya M., Hu D., Fox L.A., Dang N., Johnston E.D., Oakes J.A., Tchao B.N. Yeast life span extension by depletion of 60s ribosomal subunits is mediated by Gcn4. Cell. 2008; 133: 292-302. 10.1016/j.cell.2008.02.037. 18423200 Torgovnick A., Schiavi A., Shaik A., Kassahun H., Maglioni S., Rea S.L., Johnson T.E., Reinhardt H.C., Honnen S., Schumacher B. BRCA1 and BARD1 mediate apoptotic resistance but not longevity upon mitochondrial stress in Caenorhabditis elegans. EMBO Rep. 2018; 19: e45856. 10.15252/embr.201845856. 30366941 Rea S.L., Ventura N., Johnson T.E. Relationship between mitochondrial electron transport chain dysfunction, development, and life extension in Caenorhabditis elegans. PLoS Biol. 2007; 5e259. 10.1371/journal.pbio.0050259. 17914900 Satsuka A., Mehta K., Laimins L. p38MAPK and MK2 pathways are important for the differentiation-dependent human papillomavirus life cycle. J. Virol. 2015; 89: 1919-1924. 10.1128/JVI.02712-14. 25410865 Reinhardt H.C., Aslanian A.S., Lees J.A., Yaffe M.B. p53-Deficient Cells Rely on ATM- and ATR-Mediated Checkpoint Signaling through the p38MAPK/MK2 Pathway for Survival after DNA Damage. Cancer Cell. 2007; 11: 175-189. 10.1016/j.ccr.2006.11.024. 17292828 Manke I.A., Nguyen A., Lim D., Stewart M.Q., Elia A.E., Yaffe M.B. MAPKAP kinase-2 is a cell cycle checkpoint kinase that regulates the G2/M transition and S phase progression in response to UV irradiation. Mol. Cell. 2005; 17: 37-48. 10.1016/j.molcel.2004.11.021. 15629715 Garcia-Muse T., Boulton S.J. Distinct modes of ATR activation after replication stress and DNA double-strand breaks in Caenorhabditis elegans. EMBO J. 2005; 24: 4345-4355. 10.1038/sj.emboj.7600896 Dmitrieva N.I., Celeste A., Nussenzweig A., Burg M.B. Ku86 preserves chromatin integrity in cells adapted to high NaCl. Proc. Natl. Acad. Sci. USA. 2005; 102: 10730-10735. 10.1073/pnas.0504870102 Desler C., Lykke A., Rasmussen L.J. The effect of mitochondrial dysfunction on cytosolic nucleotide metabolism. J. Nucleic Acids. 2010; 2010: 701518. 10.4061/2010/701518 Kuchta R.D., Stengel G. Mechanism and evolution of DNA primases. Biochim. Biophys. Acta (BBA)-Bioenerg. 2010; 1804: 1180-1189. 10.1016/j.bbapap.2009.06.011 Guilliam T.A., Keen B.A., Brissett N.C., Doherty A.J. Primase-polymerases are a functionally diverse superfamily of replication and repair enzymes. Nucleic Acids Res. 2015; 43: 6651-6664. 10.1093/nar/gkv625 Zou L., Elledge S.J. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science. 2003; 300: 1542-1548. 10.1126/science.1083430. 12791985 Tsang W.Y., Lemire B.D. Mitochondrial Genome Content Is Regulated during Nematode Development. Biochem. Biophys. Res. Commun. 2002; 291: 8-16. 10.1006/bbrc.2002.6394. 11829454 Derheimer F.A., O'Hagan H.M., Krueger H.M., Hanasoge S., Paulsen M.T., Ljungman M. RPA and ATR link transcriptional stress to p53. Proc. Natl. Acad. Sci. USA. 2007; 104: 12778-12783. 10.1073/pnas.0705317104. 17616578 Hamperl S., Cimprich K.A. The contribution of co-transcriptional RNA:DNA hybrid structures to DNA damage and genome instability. DNA Repair. 2014; 19: 84-94. 10.1016/j.dnarep.2014.03.023. 24746923 Sollier J., Cimprich K.A. Breaking bad: R-loops and genome integrity. Trends Cell Biol. 2015; 25: 514-522. 10.1016/j.tcb.2015.05.003 Proudfoot N.J., Furger A., Dye M.J. Integrating mRNA processing with transcription. Cell. 2002; 108: 501-512. 10.1016/S0092-8674(02)00617-7 Heintz C., Doktor T.K., Lanjuin A., Escoubas C., Zhang Y., Weir H.J., Dutta S., Silva-Garcia C.G., Bruun G.H., Morantte I. Splicing factor 1 modulates dietary restriction and TORC1 pathway longevity in C. elegans. Nature. 2017; 541: 102-106. 10.1038/nature20789 Kuroyanagi H., Kobayashi T., Mitani S., Hagiwara M. Transgenic alternative-splicing reporters reveal tissue-specific expression profiles and regulation mechanisms in vivo. Nat. Methods. 2006; 3: 909-915. 10.1038/nmeth944 Ventura N., Rea S.L. Caenorhabditis elegans mitochondrial mutants as an investigative tool to study human neurodegenerative diseases associated with mitochondrial dysfunction. Biotechnol. J. 2007; 2: 584-595. 10.1002/biot.200600248 Mori C., Takanami T., Higashitani A. Maintenance of mitochondrial DNA by the Caenorhabditis elegans ATR checkpoint protein ATL-1. Genetics. 2008; 180: 681-686. 10.1534/genetics.108.090704 Henderson S.T., Bonafe M., Johnson T.E. daf-16 protects the nematode Caenorhabditis elegans during food deprivation. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2006; 61: 444-460. 10.1093/gerona/61.5.444. 16720740 Bloss T.A., Witze E.S., Rothman J.H. Suppression of CED-3-independent apoptosis by mitochondrial βNAC in Caenorhabditis elegans. Nature. 2003; 424: 1066-1071. 10.1038/nature01920. 12944970 Arsenovic P.T., Maldonado A.T., Colleluori V.D., Bloss T.A. Depletion of the C. elegans NAC engages the unfolded protein response, resulting in increased chaperone expression and apoptosis. PLoS ONE. 2012; 7e44038. 10.1371/journal.pone.0044038. 22957041 Kim S.K., Lund J., Kiraly M., Duke K., Jiang M., Stuart J.M., Eizinger A., Wylie B.N., Davidson G.S. A gene expression map for Caenorhabditis elegans. Science. 2001; 293: 2087-2092. 10.1126/science.1061603 Houtkooper R.H., Mouchiroud L., Ryu D., Moullan N., Katsyuba E., Knott G., Williams R.W., Auwerx J. Mitonuclear protein imbalance as a conserved longevity mechanism. Nature. 2013; 497: 451-457. 10.1038/nature12188 West A.P., Khoury-Hanold W., Staron M., Tal M.C., Pineda C.M., Lang S.M., Bestwick M., Duguay B.A., Raimundo N., MacDuff D.A. Mitochondrial DNA stress primes the antiviral innate immune response. Nature. 2015; 520: 553-557. 10.1038/nature14156 Lewis S.C., Joers P., Willcox S., Griffith J.D., Jacobs H.T., Hyman B.C. A rolling circle replication mechanism produces multimeric lariats of mitochondrial DNA in Caenorhabditis elegans. PLoS Genet. 2015; 11e1004985. 10.1371/journal.pgen.1004985 Starokadomskyy P., Gemelli T., Rios J.J., Xing C., Wang R.C., Li H., Pokatayev V., Dozmorov I., Khan S., Miyata N. DNA polymerase-alpha regulates the activation of type I interferons through cytosolic RNA:DNA synthesis. Nat. Immunol. 2016; 17: 495-504. 10.1038/ni.3409 Milenkovic D., Blaza J.N., Larsson N.G., Hirst J. The Enigma of the Respiratory Chain Supercomplex. Cell Metab. 2017; 25: 765-776. 10.1016/j.cmet.2017.03.009 George R., Beddoe T., Landl K., Lithgow T. The yeast nascent polypeptide-associated complex initiates protein targeting to mitochondria in vivo. Proc. Natl. Acad. Sci. USA. 1998; 95: 2296-2301. 10.1073/pnas.95.5.2296. 9482879 George R., Walsh P., Beddoe T., Lithgow T. The nascent polypeptide-associated complex (NAC) promotes interaction of ribosomes with the mitochondrial surface in vivo. FEBS Lett. 2002; 516: 213-216. 10.1016/S0014-5793(02)02528-0 Yogev O., Karniely S., Pines O. Translation-coupled translocation of yeast fumarase into mitochondria in vivo. J. Biol. Chem. 2007; 282: 29222-29229. 10.1074/jbc.M704201200. 17666392 Yang K.-S., Kim H.-S., Jin U.-H., Lee S., Park J.-A., Lim Y., Pai H.-S. Silencing of NbBTF3 results in developmental defects and disturbed gene expression in chloroplasts and mitochondria of higher plants. Planta. 2007; 225: 1459-1469. 10.1007/s00425-006-0453-3 del Alamo M., Hogan D.J., Pechmann S., Albanese V., Brown P.O., Frydman J. Defining the specificity of cotranslationally acting chaperones by systematic analysis of mRNAs associated with ribosome-nascent chain complexes. PLoS Biol. 2011; 9e1001100. 10.1371/journal.pbio.1001100 Amanchy R., Periaswamy B., Mathivanan S., Reddy R., Tattikota S.G., Pandey A. A curated compendium of phosphorylation motifs. Nat. Biotechnol. 2007; 25: 285-286. 10.1038/nbt0307-285 Johnson S.C., Yanos M.E., Kayser E.B., Quintana A., Sangesland M., Castanza A., Uhde L., Hui J., Wall V.Z., Gagnidze A. mTOR inhibition alleviates mitochondrial disease in a mouse model of Leigh syndrome. Science. 2013; 342: 1524-1528. 10.1126/science.1244360 Zid B.M., Rogers A.N., Katewa S.D., Vargas M.A., Kolipinski M.C., Lu T.A., Benzer S., Kapahi P. 4E-BP extends lifespan upon dietary restriction by enhancing mitochondrial activity in Drosophila. Cell. 2009; 139: 149-160. 10.1016/j.cell.2009.07.034 Bonawitz N.D., Chatenay-Lapointe M., Pan Y., Shadel G.S. Reduced TOR Signaling Extends Chronological Life Span via Increased Respiration and Upregulation of Mitochondrial Gene Expression. Cell Metab. 2007; 5: 265-277. 10.1016/j.cmet.2007.02.009 Garelick M.G., Mackay V.L., Yanagida A., Academia E.C., Schreiber K.H., Ladiges W.C., Kennedy B.K. Chronic rapamycin treatment or lack of S6K1 does not reduce ribosome activity in vivo. Cell Cycle. 2013; 12: 2493-2504. 10.4161/cc.25512 Jones M.R., Huang J.C., Chua S.Y., Baillie D.L., Rose A.M. The atm-1 gene is required for genome stability in Caenorhabditis elegans. Mol. Genet. Genom. 2012; 287: 325-335. 10.1007/s00438-012-0681-0 Kamath R.S., Ahringer J. Genome-wide RNAi screening in Caenorhabditis elegans. Methods. 2003; 30: 313-321. 10.1016/S1046-2023(03)00050-1 Butler J.A., Ventura N., Johnson T.E., Rea S.L. Long-lived mitochondrial (Mit) mutants of Caenorhabditis elegans utilize a novel metabolism. FASEB J. 2010; 24: 4977-4988. 10.1096/fj.10-162941. 20732954 Schneider C.A., Rasband W.S., Eliceiri K.W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods. 2012; 9: 671-675. 10.1038/nmeth.2089. 22930834 Pothof J., van Haaften G., Thijssen K., Kamath R.S., Fraser A.G., Ahringer J., Plasterk R.H., Tijsterman M. Identification of genes that protect the C. elegans genome against mutations by genome-wide RNAi. Genes Dev. 2003; 17: 443-448. 10.1101/gad.1060703 Wu Y.-C., Stanfield G.M., Horvitz H.R. NUC-1, a Caenorhabditis elegans DNase II homolog, functions in an intermediate step of DNA degradation during apoptosis. Genes Dev. 2000; 14: 536-548. 10.1101/gad.14.5.536. 10716942 Bhaskaran S., Butler J.A., Becerra S., Fassio V., Girotti M., Rea S.L. Breaking Caenorhabditis elegans the easy way using the Balch homogenizer: An old tool for a new application. Anal. Biochem. 2011; 413: 123-132. 10.1016/j.ab.2011.02.029 Huang D., Zhang Y., Chen X. Analysis of intracellular nucleoside triphosphate levels in normal and tumor cell lines by high-performance liquid chromatography. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2003; 784: 101-109. 10.1016/S1570-0232(02)00780-8 Hoaglin D.C., Mosteller F., Tukey J.W.. Understanding Robust and Exploratory Data Analysis; Wiley: New York, NY, USA. 1983 Dupont W.D., Plummer W.D. Jr. Power and sample size calculations. A review and computer program. Control. Clin. Trials. 1990; 11: 116-128. 10.1016/0197-2456(90)90005-M Dupont W.D., Plummer W.D. Jr. Power and sample size calculations for studies involving linear regression. Control. Clin. Trials. 1998; 19: 589-601. 10.1016/S0197-2456(98)00037-3

By Megan B. Borror; Milena Girotti; Adwitiya Kar; Meghan K. Cain; Xiaoli Gao; Vivian L. MacKay; Brent Herron; Shylesh Bhaskaran; Sandra Becerra; Nathan Novy; Natascia Ventura; Thomas E. Johnson; Brian K. Kennedy and Shane L. Rea

Reported by Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author

Titel:
Inhibition of ATR Reverses a Mitochondrial Respiratory Insufficiency
Autor/in / Beteiligte Person: Borror, Megan B. ; Girotti, Milena ; Kar, Adwitiya ; Cain, Meghan K. ; Gao, Xiaoli ; MacKay, Vivian L. ; Herron, Brent ; Bhaskaran, Shylesh ; Becerra, Sandra ; Novy, Nathan ; Ventura, Natascia ; Johnson, Thomas E. ; Kennedy, Brian K. ; Rea, Shane L.
Link:
Zeitschrift: Cells, Jg. 11 (2022-05-01), Heft 11, S. 1731-1731
Veröffentlichung: MDPI AG, 2022
Medientyp: academicJournal
ISSN: 2073-4409 (print)
DOI: 10.3390/cells11111731
Schlagwort:
  • aging
  • ageing
  • checkpoint response
  • DNA damage response
  • DDR
  • MAK-1
  • Cytology
  • QH573-671
Sonstiges:
  • Nachgewiesen in: Directory of Open Access Journals
  • Sprachen: English
  • Collection: LCC:Cytology
  • Document Type: article
  • File Description: electronic resource
  • Language: English

Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

oder
oder

Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

oder
oder

Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

xs 0 - 576
sm 576 - 768
md 768 - 992
lg 992 - 1200
xl 1200 - 1366
xxl 1366 -