New & Noteworthy

Species Can’t Risk the New Coke

January 29, 2015


Genome organization may protect key genes from the ravages of increased mutation rate during meiosis:

Back in 1985, Coca Cola decided to completely rejigger the flavor of their flagship soft drink, calling it the New Coke. This radical change to the product was a colossal failure. Toying with such an essential part of a key product was simply too risky a move. If only they had learned from our favorite beast, Saccharomyces cerevisiae.

If only Coke had protected its essential recipe as well as yeast protects its essential genes! Image via Wikimedia Commons

In a new study in PLOS Genetics, Rattray and coworkers show that the mutation rate is higher during meiosis in yeast because of the double-strand breaks associated with recombination. This makes sense, because any new mutations need to be passed on to the next generation for evolution to happen, and germ cells are made by meiosis. But their results also bring up the possibility that key genes might be protected from too many mutations by being in recombination cold spots. Unlike the Coca Cola company, yeast (and everything else) may protect essential genes from radical change.

Previous work in the Strathern lab had suggested that when double strand breaks (DSBs) in the DNA are repaired, one result is an increased mutation rate in the vicinity. The major culprit responsible for the mutations appeared to be DNA polymerase zeta (Rev3p and Rev7p).

To test whether the same is true for the DSBs that happen during the first meiotic prophase, Rattray and coworkers created a strain that contained the CAN1 gene linked to the HIS3 gene. The idea is that mutants in the CAN1 gene can be identified as they will be resistant to canavanine. The HIS3 gene is included as a way to rule out yeast that have become canavanine resistant through a loss of the CAN1 gene. So the authors were looking for strains that were both resistant to canavanine and could grow in the absence of histidine.

The first things the authors found was that the mutation rate during meiosis was indeed increased as compared to mitosis in diploids. For example, when the reporter cassette was inserted into the BUD5 gene, the mitotic mutation rate was 5.7 X 10-8 while the meiotic mutation rate was 3.7 X 10-7, a difference of around 6.5 fold.

This effect was dependent on the DSBs associated with recombination, since the increased mutation rate wasn’t seen in a spo11 mutant; the SPO11 gene is required for these breaks. Using a rev3 mutant, the authors could also conclude that at least half of the increased mutation rate is due to DNA polymerase zeta. This all strongly suggests that the act of recombination increases the local mutation rate.

If recombination is associated with the mutation rate, then areas on the genome that recombine more frequently should have a higher rate of mutation during meiosis. And they do. The authors inserted their cassette into a known recombination hotspot between the BUD23 and the ARE1 genes and saw a meiotic mutation rate of 1.77 X 10-6  as compared to a rate of 4.9 X 10-7 when inserted into a recombination coldspot. This 3.6 fold increase provides additional evidence that recombination is an important factor in meiotic recombination.

This may be more than just an unavoidable side effect of recombination. It could be that yeast and perhaps other beasts end up with their genes arrayed in such a way as to protect important genes by placing them in recombination dead zones.

And perhaps genes where lots of variation is tolerated or even helpful are placed in active recombination areas. In keeping with this, recent studies have shown that essential S. cerevisiae genes tend to be located in recombination cold spots, and that this arrangement is conserved in other yeasts.

It is too early to tell yet how pervasive this sort of gene placement is.  But if this turns out to be a good way to protect essential genes, Coca Cola should definitely have left the Coke formula in a part of its genome with little or no recombination. Mutating that set of instructions was as disastrous as mutating an essential gene!

by D. Barry Starr, Ph.D., Director of Outreach Activities, Stanford Genetics

Not Lost Without Translation

January 22, 2015


Sometimes, important information gets lost during translation. But new research shows that translation isn’t even necessary for adding a certain kind of sequence information to proteins. Image by Michael Cote via Flickr

We all learned in biology class that nucleotides get added to an mRNA using a DNA template. And that amino acids get added to proteins using an mRNA template. But as with most everything in biology, there are exceptions.

For example, a long string of A’s gets added to mRNAs in eukaryotes without a DNA template of T’s. And now, in a new study published in Science, Shen and co-workers have shown that in certain cases amino acids can be added to proteins without translating an mRNA template.

Specifically, these authors showed that threonines and alanines that are not encoded by mRNA can be added to polypeptide chains stalled on the ribosome and that a key protein in this process is Rqc2p. It makes sense that Rqcp is on the spot to do this job, as this protein is part of the ribosome quality control (RQC) complex whose job it is to ubiquinate proteins stalled at the ribosome to target them for destruction.    

These added amino acids aren’t the result of some glitch of a misbehaving cell. They appear to be critical for the cells to mount a response to a situation where ribosomes are failing to complete normal translation.

Working with our favorite model organism, S. cerevisiae, the researchers started out to investigate the ribosome quality control (RQC) complex. This complex is like a cleanup crew for stalled ribosomes. If something goes wrong during translation and the ribosome stops elongating the nascent protein chain, the RQC complex steps in and tags the partially synthesized protein with ubiquitin, marking it for degradation.

The scientists were hoping to figure out how the RQC complex finds and recognizes stalled ribosomes. So they immunoprecipitated RQC and used cryo-electron microscopy to look at the structure of the complex bound to stalled ribosomes.

After a ribosome stops translating, it falls apart into its large and small subunits. Shen and colleagues found that one component of the RQC complex, Rqc2p, binds to the large (60S) ribosomal subunit after dissociation. They also found something unanticipated: tRNAs were present in the 60S ribosomal subunit at the A and P sites. This is where the tRNAs normally reside during translation, but translation obviously couldn’t be happening, since there was no mRNA present.

The presence of a tRNA at the ribosomal A site was especially surprising because tRNAs don’t bind there stably; they need mRNA and elongation factors to stabilize the interaction. It turned out that what was keeping the tRNAs on the ribosomal large subunit was Rqc2p, which bound to both of them and stabilized them. The researchers used a new thermostable reverse transcriptase and deep sequencing to find that the bound tRNAs were two specific alanine and threonine tRNA species. Why these specific tRNAs?

Pursuing this question, Shen and colleagues made another unexpected discovery: nascent polypeptide chains from stalled ribosomes were smaller in the rqc2 null mutant than in wild type.  This suggested that Rqc2p was adding something extra to the unfinished, stalled proteins.

Putting these observations together, the researchers formulated the hypothesis that Rqc2p mediates the addition of extra alanine and threonine residues to the C termini of proteins whose translation has been stalled. They created an ingenious set of reporter constructs to test this hypothesis.

The basic reporter contained the green fluorescent protein (GFP) gene fused to a coding sequence containing multiple “difficult” codons that would cause the ribosome to stall. The researchers found that most of the strains that were mutant for different subunits of the RQC complex contained a smear of variably sized protein products, the size of GFP and larger. But the rqc2 mutant only contained unmodified GFP; it failed to add anything. This confirmed the earlier suggestion that Rqc2p was responsible for adding the mysterious extra mass.

Next, they added a protease cleavage site at various locations in the reporter gene, and found that the extra mass was added at or downstream of the stalling sequence. In other words, it was added to the C terminus of the nascent polypeptide.

To be completely sure that translation was not involved, the researchers put stop codons in every frame after the stalling sequence. They had no effect, so the extra mass couldn’t be attributed to translation in any frame.

Finally, the scientists analyzed the C-terminal extensions by total amino acid analysis, Edman degradation, and mass spectrometry. They found that the extensions consisted of between 5 and 19 alanine and threonine residues, in no defined sequence. They named them Carboxy-terminal Ala and Thr extensions, or CAT tails.

This is all very cool, but do the tails actually do anything? Yes, it looks like they do!

When translation stalls, the cell responds to this stress using the transcription factor Hsf1p. By mutating three conserved residues in the Rqc2p NFACT nucleotide-binding domain, Shen and colleagues were able to generate a protein that could still recognize the plugged-up ribosome, but couldn’t add the alanines or threonines. It still worked fine to clean up stalled proteins, but the stalled proteins had no CAT tails. And sure enough, there was no Hsf1p-mediated heat shock response either.

So the CAT tails are part of the signal that tells the cell it had better start a stress response because things aren’t looking too good at the ribosome. It’s still not obvious exactly how the CAT tails participate in this process. But this isn’t some peculiarity of yeast: the genes are conserved, and mutations in homologs of some of the RQC genes and other genes involved in translation quality control cause neurodegeneration in mice.

The ribosome is one of the best-studied molecular machines, and you might have thought we already knew just about everything there was to know about it. This work reminds us that no matter how familiar something seems, there is always more to learn when we pay attention to unexpected results.

by Maria Costanzo, Ph.D., Senior Biocurator, SGD

How to Make A Safe and Fun Mitochondrion

January 13, 2015


Influencing mitochondrial import to treat disease:

Bounce houses are a great way for kids to burn off their excess energy. They can bounce off the floor and walls and scream to their hearts’ content.

It’s important to keep tabs on how many kids get into a bounce house, so that everyone has a good time. It’s even more important for yeast and human cells to keep tabs on mitochondrial import to ensure healthy ATP synthesis. Image via Flickr

Of course, adults need to keep an eye on how many kids are in the house at any one time, to keep things safe. And if one child starts to push and kick the others, it might be easier to restore calm if the adults are careful about how many kids, and which ones, they allow inside.

The yeast mitochondrion is actually a lot like a bounce house. It’s full of energy, and it has multiple gatekeepers—protein complexes in the mitochondrial membrane that imported proteins must pass through on their way in.

And, just like a bounce house, things can go very wrong inside the mitochondrion if its proteins don’t behave properly. The end result isn’t just an upset child with a black eye, either. Genetic diseases that affect mitochondrial function are among the most severe and the hardest to treat.

Now, described in a new paper in Nature CommunicationsAiyar and colleagues have used a yeast model of human mitochondrial disease to discover both a drug and a genetic means to regulate a mitochondrial import complex. Surprisingly, tweaking mitochondrial import slightly by either of these methods mitigated the disease symptoms in both yeast and human cells.  They found a gatekeeper who can make sure there is the right number of kids in the bounce house and that they’re all behaving properly (at least, as well as they can!).

The researchers were interested in mitochondrial disorders that affected ATP synthase. This huge molecular machine in the mitochondrial inner membrane is responsible for generating most of the cell’s energy, so if it doesn’t work properly it can be a disaster for both yeast and human cells.

Aiyar and coworkers used a genetic trick to create a yeast model that had lower amounts of functional ATP synthase. This mimics many mitochondrial disorders.

They were able to reduce the amount of functional ATP synthase by using an fmc1 null mutant. Fmc1p is involved in assembly of the complex, so the fmc1 null mutant has lower amounts of functional ATP synthase and a reduced respiration rate.

First, they looked for a drug that would mitigate the effects of the fmc1 mutation. They tested the drugs in a collection that had already been FDA approved—a drug repurposing library—to see if any would improve the mutant’s respiratory growth.

The one candidate drug that emerged from the screen was sodium pyrithione (NaPT), which is used as an antiseptic. Not only did it improve the respiration of the yeast fmc1 mutant, it also improved the respiratory growth of a human cell line carrying the atp6-T8993G mutation found in patients with neuropathy, ataxia and retinitis pigmentosa (NARP, one type of ATP synthase disorder).

Aiyar and colleagues wondered exactly what was being affected by the NaPT. To figure this out, they used the S. cerevisiae genome-wide heterozygous deletion mutant collection. This is a set of diploid strains, each heterozygous for a null mutation of a different gene, that has been an incredibly useful resource for all kinds of studies in yeast.

They tested the effect of NaPT on each of the mutant strains and found that strains with mutations in the TIM17 and TIM23 genes were among the most sensitive. And, when they checked the data from previous chemogenomic screens, they saw that these two mutants were much more sensitive to NaPT than to any other drug, showing that the effect was specific.

TIM17 and TIM23 are both subunits of the Tim23 complex in the mitochondrial inner membrane that acts as a gate for many of the proteins that end up in mitochondria. The researchers found that NaPT specifically inhibited the function of this mitochondrial gatekeeper complex in an in vitro mitochondrial import assay, confirming its selectivity.

So, Aiyar and coworkers had found a drug that alleviates the effects of an ATP synthase disorder by modulating the function of a mitochondrial gatekeeper. This in itself was a huge advance: the discovery that a potentially useful, already-approved drug has a specific effect on this disease phenotype.

However, the scientists took things a step further by looking to see whether a genetic therapy could accomplish the same thing as the drug.  It was already known that overexpressing Tim21p, a regulatory subunit of the Tim23 complex, could modulate the function of the complex similarly to the effects they had seen for NaPT.

So the researchers tested whether overexpressing Tim21p would improve respiratory growth of the fmc1 mutant. Sure enough, it did. Consistent with this, assembly of the respiratory enzyme complexes of the mitochondrial inner membrane was more efficient when Tim21p was overexpressed.

Most importantly, overexpression of Tim21p in the fmc1 mutant cells caused their total ATP synthesis to more than double.  And even more exciting was the discovery that overexpressing TIMM21, the human ortholog of TIM21, in the NARP disease human cell line improved survival of those cells.

So, just like a parent deciding how many kids should be in a bounce house so that everyone has a good time, the Tim23 complex can be made to “decide” which proteins, or perhaps how many proteins, get into mitochondria, with the end result that ATP synthesis happens as efficiently as possible. The exact mechanism of this effect is still unclear, but it is clear that modulating import in this way can improve mitochondrial health even when disease mutant proteins are present.

The next step will be to translate this discovery into therapies that will help mitochondrial disease patients. People with various mitochondrial disorders may finally be able to turn their mitochondria into safe, fun places.

by Maria Costanzo, Ph.D., Senior Biocurator, SGD

Yeast Genetics Makes Awesome Sauce

January 8, 2015


Just as many different combinations of ingredients can make a homemade tomato sauce, many different combinations of alleles can mitigate the effects of a devastating mutation. Image via Pixabay.com


What’s for dinner tonight? For many of us the answer will be “pasta with tomato sauce”, even if we don’t have Italian roots.

But as you know, there isn’t any single recipe for homemade tomato sauce. Onions and garlic, or just garlic? Pork, beef, or no meat at all? How many bay leaves go in the pot? Every cook will use a slightly different combination of ingredients, but all will end up with tomato sauce.

When it comes to combining different allele variants to survive a lethal challenge, yeast is a lot like those cooks. In a new paper in GENETICS, Sirr and colleagues used divergent yeast strains to generate a wide range of allelic backgrounds and found that there is more than one way to survive a deadly mutation in the GAL7 gene.  Just as there is more than one way to make a delicious tomato sauce…

This isn’t just an academic exercise either.  GAL7 is the ortholog of the human GALT gene, which when mutated leads to the disease galactosemia. And just like in yeast, people with different genetic backgrounds may do better or worse when both copies of their GALT gene are mutated.  

GAL7 and GALT encode an enzyme, galactose-1-phosphate uridyl transferase, that breaks down the sugar galactose. If people with this mutation eat galactose, the toxic compound galactose-1-phosphate accumulates; this can cause serious symptoms or even death. The same is often true for yeast with a mutated GAL7 gene.

Ideally we would want to be able to predict how severe the symptoms of galactosemia would be, based on a patient’s genetic background. So far, though, it’s been a challenge to identify comprehensively the whole set of genes that affect a given human phenotype. Which is why Sirr and colleagues turned to our friend S. cerevisiae to study alleles affecting the highly conserved galactose utilization pathway.

The researchers started with two very divergent yeast strains, one isolated from a canyon in Israel and the other from an oak tree in Pennsylvania. Both were able to utilize galactose normally, but the scientists made them “galactosemic” by knocking out the GAL7 gene in each.

After mating the strains to create a galactosemic diploid, the researchers needed to let the strain sporulate and isolate haploid progeny. But its sporulation efficiency wasn’t very good, only about 20%. And they needed to have millions of progeny to get a comprehensive look at genetic backgrounds.

To isolate a virtually pure population of haploid progeny, Sirr and coworkers came up with a neat trick. They added a green fluorescent protein gene to the strain and put it under control of a sporulation-specific promoter.

Cells that were undergoing sporulation would fluoresce and could be separated from the others by fluorescence-activated cell sorting (FACS). The FACS technique also allowed sorting by size, so they could select complete tetrads containing four haploid spores and discard incomplete products of meiosis such as dyads containing only two spores.

After using this step to isolate tetrads, the researchers broke them open to free the spores and put them on Petri dishes containing galactose—an amount that was enough to kill either of the parent strains. One in a thousand spores was able to survive the galactose toxicity. Recombination between the divergent alleles from the two parent strains somehow came up with the right combination of alleles for a survival sauce.

Sirr and colleagues individually genotyped 247 of the surviving progeny, using partial genome sequencing. They mapped QTLs (quantitative trait loci) to identify genomic regions associated with survival. If they found a particular allele in the survivors more often than would be expected by chance, that was a clue that a gene in that region had a role in survival.

We don’t have the space here to do justice to the details of the results, but we can summarize by saying that a whole variety of factors contributed to the galactose tolerance of the surviving progeny. They had three major QTLs, regions where multiple alleles were over- or under-represented. The QTLs were centered on genes involved in sugar metabolism: GAL3 and GAL80, both involved in transcriptional regulation of galactose utilization genes, and three hexose transporter genes (HXT3, HXT6, and HXT7) that are located very close to each other.

It makes sense that all of these genes could affect galactosemia. Gal3p and Gal80p are regulators of the pathway, so alleles of these genes that make galactose catabolism less active would result in less production of toxic intermediates. And although the hexose transporters don’t transport galactose as their preferred substrates, they may induce the pathway by allowing a little galactose into the cell. So less active alleles of these transporters would also result in less galactose catabolism.

Another event that occurred in over half of the surviving progeny was aneuploidy (altered chromosome number), most often an extra copy of chromosome XIII where the GAL80 gene is located. The same three QTL peaks were also seen in the disomic strains, though, leading the authors to conclude that the extra chromosome alone was not sufficient for survival of galactosemia.

And finally, some rare non-genetic events contributed to survival of the progeny. The authors discovered this when they found that the galactose tolerance of some of the progeny wasn’t stably inherited. This could result from differences in protein levels between individual cells. For example, if one cell happened to have lower levels of a galactose transporter than other cells, it might be more resistant to galactose.

The take-home message here is that there are many different ways to get to the same phenotype. The new method that they developed allowed the researchers to see rare combinations of alleles in large numbers of individual progeny, in contrast to other genotyping methods where progeny are pooled and only the average can be detected.

For any disease or trait the ultimate goal is to identify all the alleles of all the genes that influence it. Imagine the impact on human health, if we could look at a person’s genotype and accurately predict their phenotype!

So far, it’s been a challenge to identify these large sets of human genes in a comprehensive way. But this approach using yeast could provide a feast of data to help us understand monogenic diseases like galactosemia, cystic fibrosis, porphyria, and many more, and maybe even more complex traits and diseases. Now that’s an appetizing prospect for human disease researchers. Buon appetito!

by Maria Costanzo, Ph.D., Senior Biocurator, SGD