BAPGXII Saturday May 30, 2015

logo with APG-p19lu5oeag1iikbhs1s351gbf18j8Stanford is hosting the 12th Bay Area Population Genomics (BAPG) meeting. The Bay Area Population Genomics meeting is a great place to (re)connect with your pop gen/genomics colleagues in the area and to present your work in a talk or a poster.

BAPGXI, held in December at UC Davis, was a great event with over 100 participants and a line up of excellent talks. Thanks to the Coop lab! You can read more here, including the storified tweets. We are excited to continue this success at Stanford!

Logistics

UPDATE: Click here for detailed event program.

The meeting will take place on May 30th on the Stanford campus in the Alway building, room M106. We start at 8:30AM with breakfast and registration, Dr. Dmitri Petrov’s opening remarks will begin at 9:25am, and the first talk will be at 9:30am. The last talk (Dr. Jonathan Pritchard’s keynote) ends at 2:10pm, followed by a poster session with amazing wine, beer, and cheese! Here is a general outline of the agenda, to help you plan your day:

Breakfast and Registration in Alway Courtyard 8:30-9:25am (pick up your BAPGXII gift!)
Opening Remarks 9:25-9:30am
Talk Session 1 9:30-10:30am (20 mins per talk)
Coffee Break in Courtyard 10:30-11am
Talk Session 2 11am-12pm (20 mins per talk)
Lunch in Courtyard 12-1pm
Talk Session 3 and Keynote 1-2:10pm (2 20 min talks and 1 30 min talk)
Poster Session with Wine, Beer, and Cheese Reception at 2:10pm, ends at 3pm

Talks and Posters

Sorry. Speaker and poster slots are now full. No longer accepting sign-ups.

How to Attend BAPGXII

1. Please register here by 10am Friday, May 29th to join us at BAPGXII. Registration is free and open to all, but required.

2. Encourage your colleagues to sign up! Forward this email to your lab mailing list and watch for updates on the CEHG Facebook page and on Twitter @StanfordCEHG. Help us get the momentum going by tweeting us using #BAPGXII.

3. And finally, once you’ve signed up, all you need to do is get up early and ride-share, VTA/Caltrain or bike to our beautiful campus on May 30th. Come for the science, stay for the social! Use the Stanford campus map and this Google Map to find the Alway Building, located at 500 Pasteur Drive, Stanford, CA. Be green and consider ride-sharing: there is a dedicated tab for making travel plans in the sign up doc!

We hope to see you at Stanford!

The BAPGXII organizing committee: Bridget Algee-Hewitt (@BridgetAH), David Enard (@DavidEnard), Katie Kanagawa (@KatieKanagawa), Alison Nguyen, Dmitri Petrov (@PetrovADmitri), Susanne Tilk, and Elena Yujuico. If you have any questions, feel free to contact Bridget Algee-Hewitt at bridgeta@stanford.edu.

To follow BAPGXII on twitter, check out the hashtag: #BAPGXII and also follow @StanfordCEHG .

Advertisement

Fast Algorithm Infers Relationships in High-Dimensional Datasets

Post author Henry Li is a graduate student in the Wong Lab.

Post author Henry Li is a graduate student in the Wong Lab.

New research harnesses the powers of single value decomposition (SVD) and sparse learning to tackle the problem of inferring relationships between predictors and responses in large-scale, high-dimensional datasets.

Addressing problems in computation speed, assumptions of scarcity, and algorithm sensitivity

One major challenge that statisticians face when inferring relationships is that modern data is big and the underlying true relationships between predictors and responses are sparse and multilayered. To quickly establish connections in these datasets, Ma et al. utilize a combination of SVD and sparse learning, called thresholding SVD (T-SVD). This new algorithm solves many issues that plagued the Statistics and Big Data communities, such as the problems of computation speed, the assumption of sparcity, and the sensitivity of the algorithm to positive results. In their simulation study, T-SVD is shown to be better in relation to speed and sensitivity than existing methods such as the sequential extraction algorithm (SEA) and the iterative exclusive extraction algorithm (IEEA). As a result, the multilayered relationships between predictors and responses, which come in the form of multidimensional matrices, can be learned quickly and accurately.

Uncovering new regulatory networks

Demonstrating the application of T-SVD, Ma et al. showed that new biological insights can be gained from using T-SVD to analyze datasets from The Cancer Genome Atlas consortium. The authors focused on the ovarian cancer gene expression datasets, in which the sample size is much smaller than the number of regulators and responses measured in the study. As in a typical genomic experiment, tens of thousands of genes were probed for their expression levels; from pathway studies, we know that very few of these genes form control switches that govern the expression levels for the rest of the genome. Ma et al. inferred two different relationships, based on microRNA (miRNA) or long noncoding RNA (lncRNA). The authors showed that these regulatory relationships specifically match established cancer pathways very well. Geneticists now have two new regulatory networks to mine for understanding the roles of miRNAs and lncRNAs.

In short, T-SVD is an exciting algorithm that pushes the Statistics field forward by offering a new lens to look at large-scale multidimensional datasets. With this approach, statisticians and users of statistics, like geneticists, can gain new insights into existing datasets and tackle new research problems.

References

Ma, Xin, Luo Xiao, Wing Hung Wong. Learning regulatory programs by threshold SVD regression. Proc Natl Acad SCI USA. 2014 Nov 4; 111 (44). DOI 10.1073/pnas.1417808111

Paper author, Xin (Maria) Ma is a research associate in the Wong Lab.

Paper author, Xin (Maria) Ma is a research associate in the Wong Lab.

Afterword: CEHG Genetics and Society Symposium 2015

CEHG_Logo_Mono_Black

Founded in 2012, CEHG is a research program that fosters interdisciplinary research. Home to more than 25 faculty and more than 200 grads and postdocs, CEHG bridges the divides between various member labs across Stanford campus.

The 2015 CEHG Genetics and Society Symposium (GSS15), which took place on April 13th and 14th in Stanford’s Paul Brest Hall, was a smashing success. It featured 25 speakers from Stanford campus and the San Francisco Bay academic and scientific industry communities. Approximately 175 Stanford affiliates and non-affiliates came together to celebrate the Center’s spirit of interdisciplinary collaboration and meet with experts in the fields of computational, evolutionary and human genomics This is a significant increase from last year’s 150 attendees!

The Mission:

The Genetics and Society Symposium is integral to CEHG’s mission: it provides postdocs and graduate fellows with the opportunity to share their developing research with faculty advisors and their colleagues, encourages conversation between faculty working in diverse scientific disciplines across campus, and introduces CEHG members to speakers from around the Bay Area and beyond (and vice versa).

The Venue:

As you can see from our photos of the space and catering service, Paul Brest Hall was the perfect home for this year’s two-day symposium. The hall was spacious, the food delicious, the staff hands on, and the outdoor picnic area well suited for our lunch and coffee breaks. We enjoyed the venue so much, in fact, that CEHG staff are currently in the process of booking the space for next year!

The Speakers:

GSS15 featured four brilliant keynote speakers, each distinguished in his/her field of research.

Gene Myers and CEHG Exec Committee members Marc Feldman, Chiara Sabatti, and Carlos Bustamante

Gene Myers and CEHG Exec Committee members Marc Feldman, Chiara Sabatti, and Carlos Bustamante

Founding director of a new Systems Biology Center at the Max-Planck Institute of Molecular Cell Biology and Genetics, Dr. Eugene (Gene) Myers presented his open-sourced research on the resurrection of de novo DNA sequencing. Best known for the development of BLAST, the most widely used tool in bioinformatics and the assembler he developed at Celera that delivered the fly, human, and mouse genomes in a three-year period, Dr. Myers participated in GSS15, courtesy of DNAnexus. Follow his blog: https://github.com/thegenemyers.

Co-founding director Carlos Bustamante and Ed Green catch up during a break at GSS15.

Co-founding director Carlos Bustamante and Ed Green catch up during a break at GSS15.

Assistant Professor in Biomolecular Engineering at the University of California, Santa Cruz, Richard (Ed) Green presented his research on a novel approach for highly contiguous genome assemblies, which draws on his work as an NSF Fellow at the Max Planck Institute in Leipzig, Germany and head of an analysis consortium responsible for publishing the draft genome sequence of Neanderthal. Click here for his 2014 CARTA talk, “The Genetics of Humanness: The Neanderthal and Denisovan Genomes.

Dr. Michelle Mello, Stanford Law School and School of Medicine

Dr. Michelle Mello, Stanford Law School and School of Medicine

Michelle Mello, Professor of Law at Stanford Law School and Professor of Health Research and Policy in Stanford’s School of Medicine, presented findings from her extensive research on the ethics of data sharing. As the author of more than 140 articles and book chapters on the medical malpractice system, medical errors and patient safety, public health law, research ethics, the obesity epidemic, and pharmaceuticals, Dr. Mello provided a valuable perspective from the intersections of law, ethics, and health policy. Click here to read Dr. Mello’s SLS profile.

Dr. Ami Bhatt, Stanford Medicine

Dr. Ami Bhatt, Stanford Medicine

Ami Bhatt shared her passion for improving outcomes for patients with hematological malignancies in her talk, “Bugs, drugs, and cancer.” Best known for her recent work demonstrating the discovery of a novel bacterium using sequence-based analysis of a diseased human tissue, her research has been presented nationally and internationally and published in 2013 in the New England Journal of Medicine. Click here for links to Dr. Bhatt’s CAP profile and lab homepage.

 

We had a large group of CEHG faculty members at this year’s event, showcasing the cutting edge research being done in CEHG labs across Stanford campus and indicating considerable faculty commitment to ensuring the Center’s continuing success.

Our symposium would not be complete without our invited CEHG Fellows. These speakers were nominated by organizing committee members to present on topics relating to their CEHG-funded research projects. These young scholars embody CEHG’s continuing commitment to provide funding support to researchers as they transition from graduate studies to postdoctoral scholarships.

The Workshop:

There was standing room only when facilitators Chiara Sabatti (Associate Professor of Health Research and Policy at Stanford), Ken Lange (Chair of the Human Genetics Department at UCLA), and Suyash Shringarpure (postdoctoral scholar in Stanford’s Bustamante Lab) presented their approaches to contemporary problems in statistical genetics!

Social Media:

Did you know? CEHG is on social media!

GSS15 social media moderators, Bridget Algee-Hewitt, Jeremy Hsu, Katie Kanagawa, and Rajiv McCoy were posting live throughout both days of the event. And our efforts to reach the larger community paid off, with a total reach of 815 on Facebook and more than 7,000 impressions on Twitter!

To catch up on our GSS15 coverage, check out our Facebook page at https://www.facebook.com/StanfordCEHG?ref=hl and our Twitter feed @StanfordCEHG. Follow both to make sure you are the first to know when we post CEHG-related news and announcements.

Want to know when speaker videos from the symposium will be available on CEHG’s forthcoming youtube channel? Follow us on Facebook and Twitter!

Special Thanks:

From left to right: Bridget Algee-Hewitt, Cody Sam, Yang Li, Anand Bhaskar, and Katie Kanagawa

From left to right: Bridget Algee-Hewitt, Cody Sam, Yang Li, Anand Bhaskar, and Katie Kanagawa

The GSS15 organizing committee—including Bridget Algee-Hewitt, Anand Bhaskar, Katie Kanagawa, Yang Li, and Cody Sam—would like to take this opportunity to thank CEHG Directors Carlos Bustamante and Marc Feldman, Executive Committee members Hank Greely, Dmitri Petrov, Noah Rosenberg, and Chiara Sabatti, event volunteers Alex Adams, Maude David, and Chris Gignoux, event photographer Deneb Semprum, and everyone who attended this year’s symposium.

We hope you enjoyed attending as much as we enjoyed working behind-the-scenes. We hope to see you all again at GSS16! If you are interested in volunteering for future CEHG events, please contact us at stanfordcehg@stanford.edu.

Upcoming CEHG events:

Don’t miss our popular weekly Evolgenome seminar series, which will continue through Spring term, usually on Wednesdays at noon (location varies). Lunch is always provided. Details will follow, but here is a quick overview so you can mark your calendars!

April 29: Fernando Racimo (Nielsen/Slatkin Lab)
May 6: Pleuni Pennings (UCSF)
May 20: Kelly Harkin
June 3: Sandeep Ventakaram (Petrov Lab)
June 10: Emilia Huerta-Sanchez

A framework for identifying and quantifying fitness effects across loci

Blog author Ethan Jewett is a PhD student in the lab of Noah Rosenberg.

Blog author Ethan Jewett is a PhD student in the lab of Noah Rosenberg.

The degree to which similarities and differences among species are the result of natural selection, rather than genetic drift, is a major question in population genetics. Related questions include: what fraction of sites in the genome of a species are affected by selection? What is the distribution of the strength of selection across genomic sites, and how have selective pressures changed over time? To address these questions, we must be able to accurately identify sites in a genome that are under selection and quantify the selective pressures that act on them.

Difficulties with existing approaches for quantifying fitness effects    

A recent paper in Trends in Genetics by David Lawrie and Dmitri Petrov (Lawrie and Petrov, 2014) provides intuition about the power of existing methods for identifying genomic regions affected by purifying selection and for quantifying the selective pressures at different sites. The paper proposes a new framework for quantifying the distribution of fitness effects across a genome. This new framework is a synthesis of two existing forms of analysis – comparative genomic analyses to identify genomic regions in which the level of divergence among two or more species is smaller than expected, and analyses of the distribution of the frequencies of polymorphisms (the site frequency spectrum, or SFS) within a single species (Figure 1). Using simulations and heuristic arguments, Lawrie and Petrov demonstrate that these two forms of analysis can be combined into a framework for quantifying selective pressures that has greater power to identify selected regions and to quantify selective strengths than either approach has on its own.

Figure 1. Using the quantify the strength of purging selection. The SFS tabulates the number of polymorphisms at a given frequency in a sample of haplotypes. Under neutrality (black dots) many high-frequency polymorphisms are observed. Under purifying selection (higher values of the effective selection strength |4Nes|), a higher fraction of new mutations are deleterious, leading to fewer high-frequency polymorphisms (red and blue dots). Adapted from Lawrie and Petrov (2014).

Figure 1. Using the site frequency spectrum (SFS) to quantify the strength of purifying selection. The SFS tabulates the number of polymorphisms at a given frequency in a sample of haplotypes. Under neutrality (black dots) many high-frequency polymorphisms are observed. Under purifying selection (higher values of the effective selection strength |4Nes|), a higher fraction of new mutations are deleterious, leading to fewer high-frequency polymorphisms (red and blue dots). Adapted from Lawrie and Petrov (2014).

Lawrie and Petrov begin by discussing the strengths and weaknesses of the two existing approaches. Comparative analyses of genomic divergence are beneficial for identifying genomic regions under purifying selection, which will exhibit lower-than-expected levels of divergence among species. However, as Lawrie and Petrov note, it can be difficult to use comparative analyses to quantify the strength of selection in a region because even mild purifying selection can result in complete conservation among species within the region (Figure 2). For example, whether the population-scaled selective strength, 4Nes, in a region is 20 or 200, the same genomic signal will be observed, complete conservation.

Figure 1. Adapted from Lawrie and Petrov (2013). The evolution of several 100kb regions was simulated in 32 different mammalian species under varying strengths of selection |4Nes|. The number of substitutions in each region was then estimated using genomic evolutionary rate profiling (GERP). The plot shows the median across regions of the number of inferred substitutions. From the plot, it can be seen that, once the strength of selection exceeds a weak threshold value (3 for the example given), there is full conservation among species.

Figure 1. Adapted from Lawrie and Petrov (2013). The evolution of several 100kb regions was simulated in 32 different mammalian species under varying strengths of selection |4Nes|. The number of substitutions in each region was then estimated using genomic evolutionary rate profiling (GERP). The plot shows the median across regions of the number of inferred substitutions. From the plot, it can be seen that, once the strength of selection exceeds a weak threshold value (3 for the example given), there is full conservation among species.

In contrast to comparative approaches, analyses of within-species polymorphisms based on the site frequency spectrum (SFS) within a region can be used to more precisely quantify the strength of selection. For example, Figure 1 shows that different selective strengths can produce very different site frequency spectra. Moreover, if the SFS can be estimated precisely enough, it can allow us to distinguish between two different selective strengths (e.g., 4Nes1 = 20 and 4Nes2 = 200) that would both lead to total conservation in a comparative study, and would therefore be indistinguishable. The problem is that it takes a lot of polymorphisms to obtain an accurate estimate of the SFS, and a genomic region of interest may contain too few polymorphisms, especially if the region is under purifying selection, which decreases the apparent mutation rate. Sampling additional individuals from the same species may provide little additional information about the SFS because few novel polymorphisms may be observed in the additional sample. For example, recall that for a sample of n individuals from a wildly idealized panmictic species, the expected number of novel polymorphisms observed in the n+1st sampled individual is proportional to 1/n (Watterson1975).

A proposed paradigm

Lawrie and Petrov demonstrate that studying polymorphisms by sampling many individuals across several related species (rather than sampling more individuals within a single species) could increase the observed number of polymorphisms in a region, and therefore, could increase the power to quantify the strength of selection (Figure 3) – as long as the selective forces in the genomic region are sufficiently similar across the different species.

Figure3

Figure 3. The benefits of studying polymorphisms in many populations, rather than within a single population. Three populations (A, B, and C) diverge from an ancestral population, D. The genealogy of a single region is shown (slanted lines) with mutations in the region denoted by orange slashes. Additional lineages sampled in population A are likely to coalesce recently with other lineages (for example, the red clade in population A ) and, therefore, carry few mutations that have not already been observed in the sample. In comparison, the same number of lineages sampled from a second population are likely to carry additional independent polymorphisms (for example, the red lineages in population B). If the selective pressures at the locus in populations A and B are similar, then the SFS in the two populations should be similar, and the additional lineages in B can provide additional information about the SFS. For example, if the demographic histories and selective pressures at the locus are identical in populations A and B, and if the samples from populations A and B are sufficiently diverged, then a sample of K lineages from each population, A and B, will contain double the number of independent polymorphisms that are observed in a sample of K lineages from population A alone, providing double the number of mutations that can be used to estimate the SFS.

The need for sampling depth and breadth

Without getting bogged down in the details, it’s the rare variants that are often the most important for quantifying the effects of purifying selection, so one still has to sample deeply within each species; however, overall, sampling from additional species is a more efficient way of increasing the absolute number of variants that can be used to estimate the SFS in a region, compared with sampling more deeply within the same species.

The simulations and heuristic arguments presented by Lawrie and Petrov consider idealized cases for simplicity; however, the usefulness of approaches that consider polymorphisms across multiple species has been demonstrated in methods such as the McDonald-Kreitman test (McDonald and Kreitman, 1991), which have long been important tools for studying selection. More recent empirical applications of approaches that consider information about polymorphisms across multiple species appear to do a good job of quantifying selective pressures across genomes (Wilson et al., 2011; Gronau et al., 2013), even when species are closely related (De Maio et al., 2013). Overall, the simulations and arguments presented in Lawrie and Petrov’s paper provide useful guidelines for researchers interested in identifying and quantifying selective forces, and their recommendation to sample deeply within species and broadly across many species comes at a time when such analyses are becoming increasingly practical, given the recent availability of sequencing data from many species.

References:

  1. De Maio, N., Schlötterer, C., and Kosiol, C. (2013). Linking great apes genome evolution across time scales using polymorphism-aware phylogenetic models. Molecular biology and evolution30:2249-2262.
  2. Gronau, I., Arbiza, L., Mohammed, J., and Siepel, A. (2013). Inference of natural selection from interspersed genomic elements based on polymorphism and divergence. Molecular biology and evolution30:1159-1171.
  3. Lawrie, D.S. and Petrov, D.A. (2014). Comparative population genomics: power and principles for the inference of functionality. Trends in Genetics30:133-139.
  4. Watterson, G.A. (1975). On the number of segregating sites in genetical models without recombination. Theoretical population biology7:256-276.
  5. Wilson, D.J., Hernandez, R.D., Andolfatto, P., and Przeworski, M. (2011). A population genetics-phylogenetics approach to inferring natural selection in coding sequences. PLoS genetics7:e1002395.

Paper author: David Lawrie was a graduate student in Dmitri Petrov’s lab. He is now a postdoc at USC.

 

 

 

Testing for selection in regulatory sequences using an empirical mutational distribution

How to detect selection?

Dave Yuan is a postdoc in Dmitri Petrov's lab.

Blog author Dave Yuan is a postdoc in Dmitri Petrov’s lab.

Detecting and quantifying selection in genomes is a fundamental task of interest for evolutionary biologists. A common method for this relies on comparing patterns of polymorphism and divergence between synonymous and non-synonymous sites. Synonymous sites are expected to be almost neutral, and thus mutations at these sites are expected to be fixed or lost due to genetic drift or draft. At non-synonymous sites however, mutations may get fixed due to positive selection or lost due to purifying selection. If in a specific gene, many non-synonymous sites get fixed due to positive selection, then these sites as a group will show a high evolutionary rate. On the other hand, if in a specific gene most non-synonymous mutations are lost because of purifying selection, then these sites will show a low evolutionary rate. Importantly, to determine whether the rate is high or low, we need a group of sites that can be used as a neutral comparison. For coding regions, synonymous sites are a natural choice for this comparison. [McDonald & Kreitman 1991; Keightley & Eyre-Walker 2007, Bustamante et al. 2001].

What about non-coding sequences?

Much of the genome, however, is comprised of non-coding sequence. Such sequence may contain regulatory information critical for gene expression, the modification of which is important for phenotypic evolution. Detecting selection among regulatory variation is thus of interest to evolutionary biologists, but this has been challenging. This is because functional annotation of non-coding DNA tends to be sparse, and we currently do not understand the “regulatory genetic code.” Although selection tests developed for coding sequence have been applied to non-coding sequence [reviewed in Zhen & Andolfatto 2012], a common impediment has been the choice of a group of sites that can function as a neutral comparison. A solution to this is to generate a large number of mutations in a specific region of the genome and determine whether these mutations have functional impacts. The sites at which mutations do not appear to have function can then be used to compare other groups of sites with. In a recent paper published in Molecular Biology and Evolution, graduate students Justin Smith and Kimberly McManus and CEHG faculty Hunter Fraser describe their development and application of this novel method to test for selection among variation in mammalian regulatory elements using such null distribution of mutations.

Null distribution of random mutations

Mutagenesis technique used by Patwardhan et al. (2012) to generate a comprehensive collection of cis-regulatory element mutants and test their phenotypes in vivo (figures from Patwardhan et al., 2012)

Mutagenesis technique used by Patwardhan et al. (2012) to generate a comprehensive collection of cis-regulatory element mutants and test their phenotypes in vivo (figures from Patwardhan et al., 2012)

Generating an empirical null distribution as the neutral comparison is not a trivial task. A sufficiently large—ideally comprehensive—set of mutations needs to be engineered into the regulatory element of interest, and the mutational effects or phenotypes need to be assessed. This distribution of phenotypes is then the null distribution against which the observed variation is compared to test for selection. Fortunately, recent developments in mutagenesis coupled with high-throughput sequencing have made this possible in high-resolution. Smith et al. chose data from one such mutagenesis platform that generated over 640,000 mutant haplotypes across three mammalian enhancer sequences [Patwardhan et al. 2012]. Specifically, the library of mutant enhancers was made using polymerase cycling assembly (PCA) with oligonucleotides containing between 2-3% degeneracy. All possible single nucleotide variants of the wild-type enhancer were thus represented. The library of enhancers was then cloned into a plasmid upstream of a reporter gene along with unique identification tags. This plasmid library was both sequenced to identify the tag corresponding with the mutant enhancer and injected into mouse for in vivo reporter assay. Finally, sequencing of the cDNA from the mouse liver quantified the transcriptional abundance of the tags and hence the phenotypic effects of the mutations. For each mutation it was now clear whether it upregulated or downregulated the reporter gene or whether it had no effect.

Developing a test to compare mutations and observed variation

With this dataset, Smith et al. had a comprehensive spectrum of random mutations and their phenotypic effects as the null distribution. This allowed them to create metrics for regulatory variation that are similar to the commonly-used Ka/Ks ratio, with Ka being the rate of non-synonymous change and Ks the rate of synonymous change (no functional impact on protein and hence neutral) [Kimura 1977]. The in vivo reporter assay revealed mutations with no phenotypic impact (i.e. no change in transcriptional abundance compared to wild-type), and these are analogous to synonymous or neutral changes. The new metrics are dubbed Ku/Kn, and Kd/Kn, where Ku is the rate of change for up-regulatory mutations (those with increased expression from the in vivo reporter assay), Kd is the rate of change for down-regulatory mutations, and Kn the rate of change for mutations that didn’t change expression (silent or neutral mutations).

Metrics to compare observed mutations in the phylogeny to possible mutations seen in the mutagenesis data (Figure 1 from Smith et al 2013).

Metrics to compare observed mutations in the phylogeny to possible mutations seen in the mutagenesis data (Figure 1 from Smith et al 2013).

For their analysis, the authors chose enhancer sequences from species within the same phylogenetic orders as the mutagenized enhancers. In addition to enhancer sequences from extant species, the authors also reconstructed ancestral sequences throughout the phylogeny. Combined with the mutagenesis data, each K metric at a node in the phylogeny is then calculated as the ratio of observed (i.e. in ancestors and extant species) frequencies of silent, up-, or down-regulatory polymorphisms to the frequencies of all possible silent, up-, or down-regulatory mutations respectively. Selection is inferred by comparing the ratio of up- or down-regulatory polymorphisms to the ratio of silent mutations (i.e. Ku/Kn or Kd/Kn). A comparatively low Ku or Kd, or rate of up- or down-regulatory mutations (Ku/Kn or Kd/Kn < 1) would suggest purifying selection on the polymorphisms, while a higher rate of up- or down-regulatory mutations (Ku/Kn or Kd/Kn > 1) would suggest positive selection. Smith et al. applied their new test for selection on the three enhancers from [Patwardhan et al. 2012] across the respective phylogenetic orders: LTV1 in rodents and ALDOB and ECR11 in primates. They detected purifying selection against down-regulatory polymorphisms for all three enhancers, while positive selection for up-regulatory polymorphisms was also detected for LTV1.

Detecting selection using an empirically-derived null distribution

Making evolutionary sense of variation in the regulatory regions of the genome remains more challenging than for coding sequences. We still do not have a “neutral model of regulatory evolution” to compare observed variation against. Perhaps the most exciting element of this paper, at least for me, is the use of an empirically-derived null distribution as the neutral expectation to perform this evolutionary inquiry. Patwardhan and the Shendure group at the University of Washington had earlier published a mutagenesis technique that generated a wide spectrum of mutants [Patwardhan et al. 2009]. At this time, I was getting interested in questions on the “grammar” of gene regulation, the functional characterization of regulatory sequences, and how to understand regulatory variation evolutionarily. It was thus very exciting to see both, a massively comprehensive interrogation of the mutational consequences in a regulatory element, as well as the clever application of this data to overcome a challenging evolutionary question.

One of the strengths of the Smith et al. study is the reliance on a spectrum of random mutations as the null distribution. As the original source of all genetic variation, mutations arise in a random manner. Of those that do not exert lethal effects, they may persist by chance within a population and then eventually reach certain frequencies or even fixation under selection. Because the null distribution used by Smith et al. comprises all possible mutations, it represents the mutation spectrum prior to the actions of drift or selection. It is thus an even better neutral expectation than synonymous mutations, which may not be truly neutral. In addition, using such empirical null distribution to test for selection is not limited to just regulatory variation but can be applied to coding sequence variation to reduce bias and false signals. Furthermore, by categorizing mutational effects as up- and down-regulatory, different modes of selection acting on a regulatory element can be teased apart. The interspersion of mutations—silent, up-, or down-regulatory—across the regulatory element also reduces confounding effects of regional variation in mutation rate.

As with all science, more is hoped for the future. Towards the end of the paper, the authors discuss prospects of more high-resolution mutagenesis data and, perhaps more importantly in terms of accessibility and ease of use, ability to use limited mutagenesis to test selection with. Tissue- and organism-specificity in terms of mutational effects may also be further investigated, as well as the inclusion of mutation types other than single nucleotide substitutions (e.g. insertion/deletion, copy number variation) or consideration of genomic regional context (e.g. effect of chromatin or epistasis). Nevertheless, this study represents an exciting new method to investigate regulatory variation in evolutionary contexts, one whose development and further application I look forward to seeing.

References

Bustamante CD, Wakeley J, Sawyer S, and Hartl DL. Directional Selection and the Site-Frequency Spectrum. Genetics 159:1779-1788 (2001).

Keightley PD and Eyre-Walker A. Joint inference of the distribution of fitness effects of deleterious mutations and population demography based on nucleotide polymorphism frequencies. Genetics 177:2251-2261 (2007).

Kimura M. Preponderance of synonymous changes as evidence for the neutral theory of molecular evolution. Nature 267:275-276 (1977).

McDonald JH and Kreitman M. Adaptive Protein Evolution at the Adh Locus in Drosophila. Nature 351:652-654 (1991).

Patwardhan RP, Lee C, Litvin O, Young DL, Pe’er D, and Shendure J. High-resolution analysis of DNA regulatory elements by synthetic saturation mutagenesis. Nature Biotechnology 27:1173-1175 (2009)

Patwardhan RP, Hiatt JB, Witten DM, Kim MJ, Smith RP, May D, Lee C, Andrie JM, Lee S-I, Cooper GM, et al. Massively parallel functional dissection of mammalian enhancers in vivo. Nature Biotechnology 30:265-270 (2012).

Smith JD, McManus KF, and Fraser HB. A Novel Test for Selection on cis-Regulatory Elements Reveals Positive and Negative Selection Acting on Mammalian Transcriptional Enhancers.    Molecular Biology and Evolution 30:2509-2518 (2013).

Zhen Y and Andolfatto P. Methods to Detect Selection on Noncoding DNA in Evolutionary Genomics: Statistical and Computational Methods, Volume 2, Methods in Molecular Biology, vol. 856, edited by Anisimova M. Humana Press, New York (2012).

Paper author Justin Smith is a graduate student in Hunter Fraser's lab.

Paper author Justin Smith is a graduate student in Hunter Fraser’s lab.

A fast and accurate coalescent approximation

Blog author Suyash Shringarpure is a postdoc in Carlos Bustamante's lab.

Blog author Suyash Shringarpure is a postdoc in Carlos Bustamante’s lab. Suyash is interested in statistical and computational problems involved in the analysis of biological data.

The coalescent model is a powerful tool in the population geneticist’s toolbox. It traces the history of a sample back to its most recent common ancestor (MRCA) by looking at coalescence events between pairs of lineages. Starting from assumptions of random mating, selective neutrality, and constant population size, the coalescent uses a simple stochastic process that allows us to study properties of genealogies, such as the time to the MRCA and the length of the genealogy, analytically and through efficient simulation. Extensions to the coalescent allow us to incorporate effects of mutation, recombination, selection and demographic events in the coalescent model. A short introduction to the coalescent model can be found here and a longer, more detailed introduction can be read here.

However, coalescent analyses can be slow or suffer from numerical instability, especially for large samples. In a study published earlier this year in Theoretical Population Biology, CEHG fellow Ethan Jewett and CEHG professor Noah Rosenberg proposed fast and accurate approximations to general coalescent formulas and procedures for applying such approximations. Their work also examined the asymptotic behavior of existing coalescent approximations analytically and empirically.

Computational challenges with the coalescent

For a given sample, there are many possible genealogical histories, i.e., tree topologies and branch lengths, which are consistent with the allelic states of the sample. Analyses involving the coalescent therefore often require us to condition on a specific genealogical property and then sum over all possible genealogies that display the property, weighted by the probability of the genealogy. A genealogical property that is often conditioned on is n_t, the number of ancestral lineages in the genealogy at a time t in the past. However, computing the distribution P(n_t) of n_t is computationally expensive for large samples and can suffer from numerical instability.

A general approximation procedure for formulas conditioning on n_t

Coalescent formulas conditioning on n_t typically involve sums of the form f(x)=\sum_{n_t} f(x|n_t) \cdot P(n_t)

For large samples and recent times, these computations have two drawbacks:

–       The range of possible values for n_t may be quite large (especially if multiple populations are being analyzed) and a summation over these values may be computationally expensive.

–       Expressions for P(n_t) are susceptible to round-off errors.

Slatkin (2000) proposed an approximation to the summation in f(x) by a single term f(x|E[n_t]). This deterministic approximation was based on the observation that n_t changes almost deterministically over time, even though it is a stochastic variable in theory. Thus we can write n_t \approx E[n_t]. From Figure 2 in the paper (reproduced here), we can see that this approximation is quite accurate. The authors prove the asymptotic accuracy of this approximation and also prove that under regularity assumptions, f(x|E[n_t]) converges to f(x) uniformly in the limits of t \rightarrow 0 and t \rightarrow \infty . This is an important result since it shows that the general procedure produces a good approximation for both very recent and very ancient history of the sample. Further, the paper shows how this method can be used to approximate quantities that depend on the trajectory of n_t over time, which can be used to calculate interesting quantities such as the expected number of segregating sites in a genealogy.

Approximating E[n_t] for single populations

A difficulty with using the deterministic approximation is that E[n_t] often has no closed-form formula, and if one exists, it is typically not easy to compute when the sample is large.

For a single population with changing size, two deterministic approximations have previously been developed (one by Slatkin and Rannala 1997, Volz et al. 2009 and one by Frost and Volz, 2010, Maruvka et al., 2011). Using theoretical and empirical methods, the authors examine the asymptotic behavior and computational complexity of these approximations and a Gaussian approximation by Griffiths. A summary of their results is in the table below.

Method Accuracy
Griffith’s approximation Accurate for large samples and recent history.
Slatkin and Rannala (1997), Volz et al. (2009) Accurate for recent history and arbitrary sample size, inaccurate for very ancient history.
Frost and Volz (2010), Maruvka et al. (2011) Accurate for both recent and ancient history and for arbitrary sample size.
Jewett and Rosenberg (2014) Accurate for both recent and ancient history and arbitrary sample size, and for multiple populations with migration.

 

Approximating E[n_t] for multiple populations

Existing approaches only work for single populations of changing size and cannot account for migration between multiple populations. Ethan and Noah extend the framework for single populations to allow multiple populations with migration. The result is a system of simultaneous differential equations, one for each population. While it does not allow for analytical solutions except in very special cases, the system can be easily solved numerically for any given demographic scenario.

Significance of this work

The extension of the coalescent framework to multiple populations with migration is an important result for demographic inference. The extended framework with multiple populations allows efficient computation of demographically informative quantities such as the expected number of private alleles in a sample, divergence times between populations.

Ethan and Noah describe a general procedure that can be used to approximate coalescent formulas that involve summing over distributions conditioned on n_t or the trajectory of n_t over time. This procedure is particularly accurate for studying very recent or very ancient genealogical history.

The analysis of existing approximations to E[n_t] show that different approximations have different asymptotic behavior and computational complexities. The choice of which approximation to use is therefore often a tradeoff between the computational complexity of the approximation and the likely behavior of the approximation in the parameter ranges of interest.

Future Directions

As increasingly large genomic samples from populations with complex demographic histories become available for study, exact methods either become intractable or very slow. This work adds to a growing set of approximations to the coalescent and its extensions, joining other methods such as conditional sampling distributions and the sequentially markov coalescent. Ethan and Noah are already exploring applications of these approximate methods to reconciling gene trees with species trees. In the future, I expect that these and other approximations will be important for fast and accurate analysis of large genomic datasets.

References

[1] Jewett, E. M., & Rosenberg, N. A. (2014). Theory and applications of a deterministic approximation to the coalescent model. Theoretical population biology.

[2] Griffiths, R. C. (1984). Asymptotic line-of-descent distributions. Journal of Mathematical Biology21(1), 67-75.

[3] Frost, S. D., & Volz, E. M. (2010). Viral phylodynamics and the search for an ‘effective number of infections’. Philosophical Transactions of the Royal Society B: Biological Sciences365(1548), 1879-1890.

[4] Maruvka, Y. E., Shnerb, N. M., Bar-Yam, Y., & Wakeley, J. (2011). Recovering population parameters from a single gene genealogy: an unbiased estimator of the growth rate. Molecular biology and evolution28(5), 1617-1631.

[5] Slatkin, M., & Rannala, B. (1997). Estimating the age of alleles by use of intraallelic variability. American journal of human genetics60(2), 447.

[6] Slatkin, M. (2000). Allele age and a test for selection on rare alleles.Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences355(1403), 1663-1668.

[7] Volz, E. M., Pond, S. L. K., Ward, M. J., Brown, A. J. L., & Frost, S. D. (2009). Phylodynamics of infectious disease epidemics. Genetics183(4), 1421-1430.

Paper author Ethan Jewett is a PhD student in the lab of Noah Rosenberg.

Paper author Ethan Jewett is a PhD student in the lab of Noah Rosenberg.

Caught in the act: how drug-resistance mutations sweep through populations of HIV

Blog author Meredith Carpenter is a postdoc in Carlos Bustamante's lab.

Blog author Meredith Carpenter is a postdoc in Carlos Bustamante’s lab.

It has been over 30 years since the emergence of HIV/AIDS, yet the disease continues to kill over one million people worldwide per year [UNAIDS report]. One of the reasons that this epidemic has been so difficult to control is because HIV evolves quickly—it has a short replication time and a high mutation rate, so viruses harboring new mutations that confer drug resistance tend to arise often and spread quickly.

However, the likelihood of one of these beneficial mutations popping up and subsequently “sweeping” through the viral population—i.e., becoming more common because of the survival advantage—also depends on the underlying population genetics, much of which is still poorly understood. In a paper just published in PLoS Genetics, Pleuni Pennings, postdoc in the Petrov lab, and colleagues Sergey Kryazhimskiy and John Wakeley from Harvard tracked the genetic diversity in adapting populations of HIV to better understand how and when new mutations arise.

Mutations and populations

Mutations are usually caused by either DNA damage (e.g., from environmental factors like UV radiation) or by a mistake during DNA replication. Because HIV is a retrovirus, meaning it must copy its RNA genome into DNA before it can be reproduced in the host cell, it is especially prone to errors that happen during the replication process. The rate that these errors occur, also called the mutation rate, is constant on a per-virus basis —for example, a specific mutation might happen in one virus in a million. As a consequence, the overall number of viruses in the population determines how many new mutations will be present, with a larger population harboring more mutations at any given time.

Whether these mutations will survive, however, is related to what population geneticists call the “effective population size” (also known as Ne), which takes into account genetic diversity. Due to a combination of factors, including the purely random destruction of some viruses, not all mutations will be preserved in the population, regardless of how beneficial they are. The Ne is a purely theoretical measure that can tell us how easily and quickly a new mutation can spread throughout a population. Because it accounts for factors that affect diversity, it is usually smaller than the actual (or “census”) population size.

Pennings and colleagues wanted to determine the Ne for HIV in a typical patient undergoing drug treatment. This is a contentious area: previous researchers examining this question using different methods, including simply summing up overall mutation numbers, came up with estimates of Ne ranging from one thousand to one million (in contrast, the actual number of virus-producing cells in the body is closer to one hundred million, but more on that later). To get a more exact estimate, Pennings took a new approach. Using previously published DNA sequences of HIV sampled from patients over the course of a drug treatment regimen, she looked at the actual dynamics of the development of drug-resistant virus populations over time.

Swept away

Specifically, Pennings focused on selective sweeps, wherein an advantageous mutation appears and then rises in frequency in the population. Features of these sweeps can give estimates of Ne because they reveal information about the diversity present in the initial population. Pennings sought to distinguish between “hard” and “soft” selective sweeps occurring as the viruses became drug resistant. A hard sweep occurs when a mutation appears in one virus and then rises in frequency, whereas a soft sweep happens when multiple viruses independently gain different mutations, which again rise in frequency over time (see Figure 1). These two types of sweeps have distinct fingerprints, and their relative frequencies depend on the underlying effective population size—soft sweeps are more likely when a population is larger it becomes more likely for different beneficial mutations to independently arise in two different viruses. Soft sweeps also leave more diversity in the adapted population compared to hard sweeps (Figure 1).

Figure 1, an illustration of a hard sweep (left) and a soft sweep (right).

Figure 1, an illustration of a hard sweep (left) and a soft sweep (right).

To tell these types of sweeps apart, Pennings took advantage of a specific amino acid change in the HIV gene that encodes reverse transcriptase (RT). This change can result from two different nucleotide changes, either one of which will change the amino acid from lysine to asparagine and confer resistance to drugs that target the RT protein.  Pennings used this handy feature to identify hard and soft sweeps: if she observed both mutations in the same drug-resistant population, then the sweep was soft. If only one mutation was observed, the sweep could be soft or hard, so she also factored in diversity levels to tell these apart. Pennings found evidence of both hard and soft sweeps in her study populations. Based on the frequencies of each, she estimated the Ne of HIV in the patients. Her estimate was 150,000, which is higher than some previous estimates but still lower than the actual number of virus-infected cells in the body. Pennings suggests that this discrepancy could be due to the background effects of other mutations in the viruses that gain the drug-resistance mutation—that is, even if a virus gets the valuable resistance mutation, it might still end up disappearing from the population because it happened to harbor some other damaging mutation as well. This would reduce the effective population size as measured by selective sweeps.

Implications and future work

Pennings’ findings have several implications. The first is that HIV populations have a limited supply of resistance mutations, as evidenced by the presence of hard sweeps (which, remember, occur when a sweep starts from a single mutation). This means that even small reductions in Ne, such as those produced by combination drug therapies, could have a big impact on preventing drug resistance. The second relates to the fact that, as described above, the likelihood that a mutation will sweep the population may be affected by background mutations in the virus in which it appears. This finding suggests that mutagenic drugs, given in combination with standard antiretrovirals, could be particularly useful for reducing drug resistance.  Now, Pennings is using larger datasets to determine whether some types of drugs lead to fewer soft sweeps (presumably because they reduce Ne). She is also trying to understand why drug resistance in HIV evolves in a stepwise fashion (one mutation at a time), even if three drugs are used in combination.

Paper author Pleuni Pennings is a postdoc in the lab of Dmitri Petrov.

Paper author Pleuni Pennings is a postdoc in the lab of Dmitri Petrov.

References

Pennings, PS, Kryazhimskiy S, Wakeley J. Loss and recovery of genetic diversity in adapting HIV populations. 2014, PLoS Genetics.

Using phyloseq for the reproducible analysis of high-throughput sequencing data in microbial ecology

Blog author Diana Proctor is a graduate student in David Relman’s lab.

Blog author Diana Proctor is a graduate student in David Relman’s lab.

The Problem: Data Availability & Scientific Reproducibility

A Current Biology (1) paper evaluating the accessibility of scientific data recently inspired articles and blog posts (2, 3) as well as a lively conversation on Reddit about the “alarming rate of data disappearance” (4). Solutions to the problem of disappearing data include the NIH data sharing policy, as well as data sharing policies set by scientific journals, requiring the deposition of data into public repositories.

As a trainee in David Relman’s Lab thinking about the eventual fate of the high-throughput, next generation sequencing data generated over the course of my dissertation (http://www.hyposalivation.com), this conversation about data accessibility brings to mind a related question – how can I ensure that my data are not lost as fast as the current biology paper predicts?

The Solution: Phyloseq allows microbial ecologists to make reproducible research reports

The solution to data disappearance probably needs to involve not only deposition of data into public repositories, but also the widespread use of reproducible research reports. Luckily for those microbial ecologists among us, Paul McMurdie and Susan Holmes of Stanford University developed an R-based Bioconductor package (i.e., a package for bioinformatics) called phyloseq to facilitate the reproducible statistical analysis of high throughput phylogenetic sequencing datasets, including those generated by barcoded amplicon sequencing, metabolomic, and metagenomic experiments (5, 6). Phyloseq, initially released in 2012, was recently updated by McMurdie & Holmes, and described in an April 2013 publication (6).

Phyloseq Key Features

Phyloseq allows the user to import a species x sample data matrix (aka, an OTU Table) or data matrices from metagenomic, metabolomic, and/or other –omics type experiments into the R computing environment. Previous R extensions, such as OTUbase, also have the capacity to import these data matrices into R, but phyloseq is unique in that it allows the user to integrate the OTU Table, the phylogenetic tree, the “representative sequence” fasta file, and the metadata mapping file into a single “phyloseq-class” R object. The microbial ecologist can then harness all the statistical and graphical tools available in R, including Knitr, R-Markdown and ggplot2 (among others), to generate reproducible research reports with beautiful graphics, as detailed below. To see the report McMurdie used to prepare the phyloseq publication, visit this link: http://www.hyposalivation.com/wp-content/uploads/2014/01/phyloseq_plos1_2012-source-doc.html.

1. Phyloseq incidentally allows the user to curate data

When phyloseq imports the myriad phylogenetic sequencing data objects into R, it scrutinizes the data, making sure that the OTU Table matches the metadata mapping file, the phylogenetic tree, and the representative sequence labels. If not, the user gets an error. If the data descriptors are congruent, a single phyloseq object can be created, which can then be saved along with the R code used to create the object. I have found that this enables me to curate my data – consolidating all the data objects (OTU Table, mapping file, phylogenetic tree, etc.) describing a single experiment into a single multi-level data frame.

2. Phyloseq gives the user the analytical power of R

Importantly, by importing data into the R-computing environment, one may easily perform beta diversity analysis using any or all of over 40 different ecological distance metrics before performing virtually any ordination under the sun. Several alpha diversity metrics are implemented in phyloseq, as well. Finally, after getting the data into R, it’s easy to perform more sophisticated analyses than has previously been possible with this type of dataset, such as k-tables analysis (7), using R’s repertoire of extension packages.

3. Phyloseq makes standardization of sequence data pretty simple

Of particular note, the authors have included in phyloseq several methods to standardize and/or normalize high throughput sequence data. Most of us of course realize the need for data standardization (as evidenced by our reliance on rarefaction), but the tools to easily standardize data, aside from rarefaction, have been lacking (8). The authors of phyloseq have equipped us with several methods (one new!) to standardize our microbial census data, as well as the code needed to accomplish the task (https://github.com/joey711/phyloseq/wiki/Vignettes).

4. Phyloseq makes subsetting large datasets easy

One of my favorite uses for the phyloseq package is that it allows me to easily subset my dataset. In my work, I study the spatial variation of oral microbial communities. I have taken samples from all teeth from the mouths of just a handful of research subjects, but I have samples for certain teeth from all subjects. Phyloseq makes it easy for me to take a complete OTU Table, and subset it on only those teeth that were sampled in all subjects. Similarly, I can subset my OTU Table on a single bacterial phylum or on a single species, or on any covariate in my metadata mapping file, using a single line of R code.

5. Phyloseq enables the user to generate reproducible graphics

The authors of phyloseq created several custom ggplot2 (9) functions, enabling the phyloseq user, with just a few lines of code, to generate all of the most common graphics used in microbial census research (e.g., heatmaps, networks, ordination plots, phylogenetic trees, stacked bar plots for abundance measurements, etc.). Examples of these plots are shown in Figure 1 (though many other possibilities are supported, which can be seen here: http://joey711.github.io/phyloseq/tutorials-index).

Fig 1A. The NMDS ordination plot  shows the separation of samples by weighted UniFrac distance for the Global Patterns dataset. Human-associated communities appear to cluster towards the right side of NMDS1 while non-human associated communities cluster towards the left.

Fig 1A. The NMDS ordination plot shows the separation of samples by weighted UniFrac distance for the Global Patterns dataset. Human-associated communities appear to cluster towards the right side of NMDS1 while non-human associated communities cluster towards the left.

Fig 1B. The dodged boxplots  show three alpha diversity metrics (Observed species, Chao1, and ACE) on the Y-axis with data classified on the X-axis as either a human-associated or a non-human associated microbial community.  This plot shows that non-human associated communities, in general, appear to be much more diverse than human-associated communities.

Fig 1B. The dodged boxplots show three alpha diversity metrics (Observed species, Chao1, and ACE) on the Y-axis with data classified on the X-axis as either a human-associated or a non-human associated microbial community. This plot shows that non-human associated communities, in general, appear to be much more diverse than human-associated communities.

6. Phyloseq allows covariate data to be visualized with the phylogenetic tree

In particular, phyloseq solves very well the problem of visualizing the phylogenetic tree – it allows the user to project covariate data (such as sample habitat, host gender, etc.) onto the phylogenetic tree, so that relationships between microbes, microbial communities, and the habitat from which they were derived can easily be seen. As an example, the relative abundance of taxa in samples was projected onto the phylogenetic tree along with the environment from which the samples were derived along with the bacterial order in Figure 2. I’ve not seen any other application that allows similar visualizations of the tree, and bootstrapping is also supported. For additional examples, refer to the phyloseq tutorial (http://joey711.github.io/phyloseq/plot_tree-examples.html).

Fig 2: An example of using phyloseq to visualize phylogenetic trees along with covariate data using the Global Patterns dataset. In this figure, the sample type is shown in color, the shapes are bacterial Order, and the size of the shapes indicates the relative abundance of the taxon in the sample.

Fig 2: An example of using phyloseq to visualize phylogenetic trees along with covariate data using the Global Patterns dataset. In this figure, the sample type is shown in color, the shapes are bacterial Order, and the size of the shapes indicates the relative abundance of the taxon in the sample.

7. Data & code & results can be saved together to improve scientific reproducibility

One of the key features of phyloseq is that it provides researchers who have access to any system where R is able to run with a framework (i.e., R, R-markdown, Knitr & Rstudio) to perform reproducible statistical analysis of high throughput sequencing data. Using Phyloseq, Rstudio, R-markdown, and Knitr, it’s possible to see in a single .html file the data used to generate a set of figures alongside the code that was used to generate those figures. I now keep a collection of reproducible research reports as part of my lab notebook, and I look forward to being able to publish the final report for my first study along with my first scientific manuscript. For an example, please see the phyloseq tutorials, which were also generated using this approach (http://joey711.github.io/phyloseq/import-data.html).

8. Phyloseq is easy to learn.

When I first began working with Phyloseq, after taking a class taught by the author Susan Holmes, I knew some basic R commands from an undergraduate statistics class. Working with phyloseq made learning R easy for me. Since Phyloseq has a built in set of datasets one can use, it’s easy to reproduce the figures published in the phyloseq paper, as a stepping-stone for creating figures of one’s own.

Conclusion

An R-based package called Phyloseq makes it easy to analyze high throughput microbial census data, visualize the data, and perform reproducible statistical analysis. Phyloseq should facilitate conversations between researchers who publish data and the consumers of it with its emphasis on reproducible research. This should help those of us in the infancy of microbiome research ensure that our data do not disappear as quickly as Vine et al. currently predicts.

The paper was co-authored by Paul McMurdie and Professor Susan Holmes (in the picture).

The paper and the phyloseq package are co-authored by Paul McMurdie and Professor Susan Holmes (in the picture).

REFERENCES

1.         Vines TH, Albert AY, Andrew RL, Debarre F, Bock DG, Franklin MT, et al. The Availability of Research Data Declines Rapidly with Article Age. Current biology : CB. 2013. doi: 10.1016/j.cub.2013.11.014. PubMed PMID: 24361065.

2.         Stromberg J. smithsonianmag.com: 2014.]. Available from: http://blogs.smithsonianmag.com/science/2013/12/the-vast-majority-of-raw-data-from-old-scientific-studies-may-now-be-missing/.

3.         Noorden EGRV. Scientists losing data at a rapid rate. Nature News. 2014.

4.         bobmanyun. reddit.com: 2014 1-7-2014. [cited 2014]. Available from: http://www.reddit.com/r/science/comments/1tb2d3/scientific_data_is_disappearing_at_alarming_rate/.

5.         McMurdie PJ, Holmes S. Phyloseq: a bioconductor package for handling and analysis of high-throughput phylogenetic sequence data. Pac Symp Biocomput. 2012:235-46. Epub 2011/12/17. PubMed PMID: 22174279; PubMed Central PMCID: PMC3357092.

6.         McMurdie PJ, Holmes S. phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS One. 2013;8(4):e61217. doi: 10.1371/journal.pone.0061217. PubMed PMID: 23630581; PubMed Central PMCID: PMC3632530.

7.         Thioulouse J. Simultaneous analysis of a sequence of paired ecological tables: A comparison of several methods. Ann Appl Stat. 2011;5(4):2300-25. doi: 10.1214/10-AOAS372.

8.         Holmes PJMS. Waste Not, Want Not: Why Rarefying Microbiome Data is Inadmissible. ARXIV. 2013. Epub 10/2013.

9.         Wickham H. ggplot2: Elegant Graphics for Data Analysis: Springer; 2009. 213 p.

Taking studies of regulatory evolution to the next level: translation

Carlo Artieri, a postdoc in the group of Hunter Fraser, wrote this blog post. The paper is written by Carlo and Hunter.

Carlo Artieri, a postdoc in the group of Hunter Fraser, wrote this blog post. The paper is written by Carlo and Hunter.

Carlo Artieri writes about his new paper: Evolution at two levels of gene expression in yeast which is in press in Genome Research.

Understanding the molecular basis of regulatory variation within and between species has become a major focus of modern genetics. For instance, the majority of identified human disease-risk alleles lie in non-coding regions of the genome, suggesting that they affect gene regulation (Epstein 2009). Furthermore, it has been argued that regulatory changes have played a dominant role in explaining uniquely human attributes (King and Wilson 1975). However, our knowledge of gene regulatory evolution is based almost entirely on studies of mRNA levels, despite both the greater functional importance of protein abundance, and evidence that post-transcriptional regulation is pervasive. The availability of high-throughput methods for measuring mRNA abundance, coupled to the lack of comparable methods at the protein level have contributed to this focus; however, a new method known as ribosome profiling, or ‘riboprofiling’ (Ingolia et al. 2009), has enabled us to study the evolution of translation in much greater detail than was possible before. This method involves the construction of two RNA-seq libraries: one measuring mRNA abundance (the ‘mRNA’ fraction), and the second capturing the portion of the transcriptome that is actively being translated by ribosomes (the ‘Ribo’ fraction). On average, the abundance of genes within the Ribo fraction should be proportional to that of the mRNA fraction. Genes with increased translational efficiency are identified when Ribo fraction abundance is higher than that of the mRNA fraction, whereas reduced translational efficiency is inferred when the opposite is observed.

Riboprofiling of yeast hybrids

We performed riboprofiling on hybrids of two closely related species of budding yeast, Saccharomyces cerevisiae and S. paradoxus, (~5 million years diverged). In hybrids, the parental alleles at a locus share the same trans cellular environment; therefore in the absence of cis-regulatory divergence in transcription, both alleles should be expressed at equal levels. Conversely, cis-regulatory divergence will produce unequal expression of alleles (termed allele-specific expression, or ‘ASE’). Cis-regulatory divergence at the translational level is detected when ASE in the mRNA fraction does not equal that measured in the Ribo fraction, indicating independent divergence across levels. We also performed riboprofiling on the two parental strains, as differences in the expression of orthologs between parental species that cannot be explained by the allelic differences in the hybrids can be attributed to trans divergence. Therefore, by measuring differences in the magnitudes of ASE between the two riboprofiling fractions in the hybrids and the parents, we identified independent cis and trans regulatory changes in both mRNA abundance and translational efficiency.

Results

We found that both cis and trans regulatory divergence in translational efficiency is widespread, and of comparable magnitude to divergence at the mRNA level – indicating that we miss much regulatory evolution by focusing on mRNA in isolation. Moreover, we observed an overwhelming bias towards divergence in opposing parental directions, indicating that while many orthologs had higher mRNA abundance in one parent, they often showed increased translational efficiency in the other parent. This suggests that stabilizing selection acts to maintain more similar protein levels between species than would be expected by comparing mRNA abundances alone.

Translational divergence not associated with TATA boxes

Interestingly, while we confirmed the results of previous studies indicating that both cis and trans regulatory divergence at the mRNA level are associated with the presence of TATA boxes and nucleosome free regions in promoters, no such relationship was found for translational divergence, indicating that these regulatory systems have different underlying architectures.

Evidence for polygenic selection at two levels

We also searched for evidence of polygenic selection in and between both regulatory levels by applying a recently developed modification of Orr’s sign test (Orr 1998; Fraser et al. 2010; Bullard et al. 2010). Under neutral divergence, no pattern is expected with regards to the parental direction of up or down-regulating alleles among orthologs within a functional group (e.g., a pathway or multi-gene complex). However, a significant bias towards one parental lineage is evidence of lineage-specific selection. This analysis uncovered evidence of polygenic selection at both regulatory levels in a number of functional groups. In particular, genes involved in tolerance to heavy metals were enriched for reinforcing divergence in mRNA abundance and translation favoring S. cerevisiae. Increased tolerance to these metals has been observed in S. cerevisiae (Warringer et al. 2011), suggesting that domesticated yeasts have experienced a history of polygenic adaptation across regulatory levels allowing them to grow on metals such as copper.

Finally, using data from the Ribo fraction, we also uncovered multiple instances of conserved stop-codon readthrough, a mechanism via which the ribosome ‘ignores’ the canonical stop codon and produces a C-terminally extended peptide. Only two cases of C-terminal extensions have previously been observed in yeast, though in one such case, PDE2, extension of the canonical protein plays a functional role in regulating cAMP levels (Namy et al. 2002). Our data suggests that this mechanism may occur in dozens of genes, highlighting yet another post-transcriptional mechanism leading to increased proteomic diversity.

Conclusion

By applying a novel approach to a long-standing question, our analysis has revealed that post-transcriptional regulation is abundant, and likely as important as transcriptional regulation. We argue that partitioning the search for the locus of selection into the binary categories of ‘coding’ vs. ‘regulatory’ overlooks the many opportunities for selection to act at multiple regulatory levels along the path from genotype to phenotype.

References

Artieri CG, Fraser HB. 2013. Evolution at two levels of gene expression in yeast. Genome Research (in press).
Preprint on the arXiv. 

Bullard JH, Mostovoy Y, Dudoit S, Brem RB. 2010. Polygenic and directional regulatory evolution across pathways in Saccharomyces. Proc Natl Acad Sci USA 107: 5058-5063.

Epstein DJ. 2009. Cis-regulatory mutations in human disease. Brief Funct Genomic Proteomic 8: 310–316.

Fraser HB, Moses AM, Schadt EE. 2010. Evidence for widespread adaptive evolution of gene expression in budding yeast. Proc Natl Acad Sci USA 107: 2977-2982.

Ingolia NT, Ghaemmaghami S, Newman JR, Weissman JS. 2009. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324:218-223.

King MC, Wilson AC. 1975. Evolution at two levels in humans and chimpanzees. Science 188: 107-116.

Namy O, Duchateau-Nguyen G, Rousset JP. 2002. Translational readthrough of the PDE2 stop codon modulates cAMP levels in Saccharomyces cerevisiae. Mol Microbiol 43: 641-652.

Orr HA. 1998. Testing natural selection vs. genetic drift in phenotypic evolution using quantitative trait locus data. Genetics 149: 2099-2104.

Warringer J, Zörgö E, Cubillos FA, Zia A, Gjuvsland A, Simpson JT, Forsmark A, Durbin R, Omholt SW, Louis EJ, Liti G, Moses A, Blomberg A. 2011. Trait variation in yeast is defined by population history. PLoS Genet 7 :e1002111.

Which genetic variants determine histone marks?

JoeDavis

Blog author Joe Davis is a graduate student with Stephen Montgomery & Carlos Bustamante.

The wealth of genetic variation in the human genome is found not within protein-coding genes but within non-protein coding regions. This comes as no surprise given that only 1% percent of the genome codes for proteins. Until recently, efforts to determine the effects of genetic variation on trait variation and disease have focused on coding regions. Results of genome-wide association studies (GWAS), however, have shown that trait and disease associated variants are often regulatory variants such as expression quantitative trait loci (eQTLs) found in non-coding regions. These results have spurred an effort to understand the functional role of non-coding, regulatory variation. Efforts have thus far relied on characterizing the association between variants and gene expression. This association alone, however, will not reveal the complete functional mechanism by which non-coding variants influence gene expression. Recent efforts have therefore begun to characterize numerous molecular phenotypes such as transcription factor (TF) binding, histone modification, and chromatin state to determine the mechanisms by which regulatory variants affect gene expression.

One issue, four papers

In the November 8 issue of Science, three papers were published that address the role of non-coding genetic variation on TF binding, histone modifications, and chromatin state (i.e. active versus inactive enhancer status). The first study was completed by the Dermitzakis Lab at the University of Geneva. They analyzed three TFs, RNA polymerase II (Pol II), and five histone modifications using chromatin immunoprecipitation and sequencing (ChIP-Seq) in lymphoblastoid cell lines (LCLs) from two parent-child trios [1]. The second was completed by the Pritchard Lab, which has recently moved to Stanford, and the Gilad Lab at the University of Chicago. They identified genetic variants affecting variation in four histone modifications and Pol II occupancy in ten unrelated Yoruba LCLs [2]. The third study was performed by the Snyder Lab at Stanford. They characterized the genetic variation underlying changes in chromatin state using RNA-Seq and ChIP-Seq for four histone modifications and two DNA binding factors in 19 LCLs from diverse populations [3]. This work was the subject of a recent CEHG Evolgenome talk given by Maya Kasowski, the study’s first author. Finally, the fourth study, published in the November 28 issue of Nature, was performed by the Glass Lab at UCSD. They characterized the effect of natural genetic variation between two mouse strains on the binding of two TFs involved in cell differentiation (PU.1 and C/EBPα) using ChIP-Seq [4]. In this post, I will analyze primarily the work presented by the Pritchard Lab, but I strongly recommend reading all four papers to understand the challenges in characterizing non-coding variation and the methods available to do so.

Motivation

The four studies seek to answer the general question of how regulatory variation affects gene expression. They characterize diverse molecular phenotypes such as histone modifications and TF binding to understand the mechanisms of action for non-coding variants. The Pritchard Lab focused their study on four histone modifications (three active and one repressive: H3K4me3, H3K4me1, H3K27ac, and H3K27me3, respectively) and Pol II occupancy.

Table2

Histone modifications 101

Histone modifications refer to the addition of chemical groups such as methyl or acetyl to specific amino acids on the tails of histone proteins comprising the nucleosome. These chemical groups are referred to as histone marks. They can serve a wide range of functions, but in general they are associated with the accessibility of a chromatin region. For example, the tri-methylation of lysine 4 of histone 3 (H3K4me3) is associated with increased chromatin accessibility and gene activation. On the other hand, increased levels of the repressive mark H3K27me3 (tri-methylation of lysine 27 of histone 3) at promoters is associated with gene inactivation.

Histone mark levels are measured in a high-throughput manner using ChIP-Seq. Briefly, an antibody targeting the mark of interest is used to pull down modified genomic regions. These immunoprecipitated regions are then sequenced to determine which genomic segments are modified and at what level. The procedure usually requires a large number of cells (on the order of 10^7). Therefore, the modification level is, in some ways, a population level measurement. Analysis of ChIP-Seq data typically involves testing for genomic regions with more reads than expected by chance. These regions, ranging from 200bp to 1000bp or more, are referred to as peaks that represent a modification level above the genomic background. Repressive marks like H3K27me3 tend to have broad peak regions, while activating marks like H3K4me3 can have much tighter peaks.

Since modification levels represent measurements on a population of cells and histone residues can have multiple modifications, genomic regions can show evidence for multiple marks. The combinations of these marks over a region can mark the function of the region. For example, regions with high levels of H3K27ac and a high ratio of H3K4me1 to H3K4me3 can mark active enhancer regions. Until now, the variation of these marks between individuals and the genetic cause of this variation was uncharacterized. Moreover, the causal impact of these marks remains unknown. Do they alter gene expression directly or are they altered by gene regulation? Therefore, the two guiding questions for this study are:

1. What genetic variants influence histone modifications?

2. Are these modifications “a cause or a consequence of gene regulation?”

Variation in histone modifications, a real whodunit

The authors first seek to identify and characterize genetic variants that influence histone marks. They generated ChIP-Seq data for the four histone marks and Pol II in LCLs derived from ten unrelated Yoruba individuals who were previously genotyped as part of the 1000 Genomes Project. Similar studies of regulatory variants such as eQTL studies require large sample sizes to detect the effects of regulatory variants that often lie outside the gene. Unlike eQTL studies, histone marks cover fairly broad regions often encompassing causal regulatory variants. As a result, the authors can use a smaller sample size and still be confident about interrogating the effects of causal regulatory SNPs. The authors developed a statistical test that models total read depth between individuals and allelic imbalance between haplotypes within individuals to increase power to detect cis-QTLs (i.e. variants that affect histone marks and Pol II occupancy nearby in the genome). Using this method, they identified over 1200 distinct QTLs for histone marks and Pol II occupancy (FDR 20%).

The authors then analyze these histone mark and Pol II QTLs to determine the overlap of these variants with other known regulatory variants. The hypothesis is that regulatory variants that affect gene expression will have effects on diverse molecular phenotypes. Therefore, variants that influence histone marks and Pol II should show significant overlap with known regulatory variants such as eQTLs and DNase I sensitivity QTLs (dsQTLs). DNase I sensitivity is a measure of chromatin accessibility with higher sensitivity associated with higher accessibility. The Pritchard Lab mapped eQTLs and dsQTLs in a larger sample of ~75 Yoruban LCLs in two previous studies that I also recommend reading [5,6]. Their analysis revealed an enrichment of low p-values for dsQTLs and, to a lesser extent, eQTLs when tested as histone mark and Pol II QTLs. In addition, the authors observed a coordinated change in multiple molecular phenotypes at dsQTLs and eQTLs. For example, higher levels of the three histone active marks were observed at dsQTLs for the more DNase I sensitive genotype. At eQTLs, H3K4me3, H3K27ac, and Pol II levels were higher for individuals with the high expression genotype. These results show that non-coding regulatory variants impact multiple molecular phenotypes ranging from chromatin accessibility and transcription to histone modifications. The authors provide strong evidence in response to their first guiding question, namely that non-coding regulatory polymorphisms associate with variation in histone marks and Pol II.

TFs and a question of directionality

The authors then turned to addressing the questions of causality for these marks. To do so, they analyze genetic variants in TF binding sites. The main hypothesis is that regulatory variants that alter a TFBS will modify TF binding which will cause changes in histone mark and Pol II levels nearby. If this is the case, then changes in histone marks are a consequence of how strong the TF binding site is. On the other hand, if these marks were causal, polymorphisms in TF binding sites would not be expected to show strong association with changes in these marks.

To test their hypothesis, the authors examine ~11.5K TF binding sites with polymorphisms heterozygous in at least 1 of their 10 individuals. They calculate the change in position weight matrix (PWM) score between the two alleles for polymorphic TF binding sites within each individual. They then test for significant association between this change in PWM and allelic imbalance of ChIP-Seq reads at nearby heterozygous sites. The idea is that if a variant improves (or disrupts) TF binding for one allele at a TF binding site then active histone marks nearby on the same allele will increase (or decrease). Repressive histone marks (in this case H3K27me3) are expected to have the opposite response. Indeed, when they apply their test, they find a significant positive association for the active marks and a negative association for the repressive mark. This result supports the hypothesis of changes histone marks as a consequence of TF binding and gene regulation. However, this result does not rule out other possibilities. Histone marks can still play a causal role in the establishment of TF binding. In other words, the relationship between TF binding and histone marks does not have to be unidirectional. In addition, there is evidence that long non-coding RNAs may play a role in the establishment and regulation of histone marks.

dsQTLs and eQTLs, a match made on chromatin

In their final analysis, the authors examine dsQTLs that are also eQTLs. Since these variants associate with both gene expression and chromatin accessibility at distal regulatory regions (>5kb from associated TSS), the authors can assign the regulatory region to a specific gene. A variant that is both a dsQTL and an eQTL likely disrupts a distal regulatory region. In addition to disrupting the accessibility of the regulatory region, the variant also perturbs the expression of a gene influenced by the regulatory region. For example, a variant may decrease the chromatin accessibility of an enhancer region and thereby decrease the level of active histone marks for the enhancer. This decreased enhancer activity can result in decreased transcription from a nearby gene and similarly decreased active mark levels for the gene. Therefore, the hypothesis guiding this analysis is that variants influencing the histone marks of a distal regulatory region will have a coordinated effect on histone marks at genes under the control of the regulatory region. The authors examine the allelic imbalance in ChIP-Seq reads at regulatory regions and their associated transcription start sites (TSS). Indeed, the authors observe that variants that increase DNase I sensitivity have a significant positive allelic imbalance for active marks at both the regulatory region and the TSS. The opposite is true for the repressive mark. This result again emphasizes the complexity of gene regulation and the impact of non-coding variation. Not only do regulatory variants influence diverse molecular phenotypes nearby, they can direct changes at distal loci. As the authors note, this coordinated change in histone marks between distal regions possibly reflects the 3D organization of chromatin. Regulatory variants that impact chromatin looping interactions between distal regulatory regions and genes may cause changes in activity levels for both the gene and the regulatory region.

Conclusions

This paper provides clear evidence that regulatory variation has very complex impacts affecting multiple and diverse molecular phenotypes at multiple regions simultaneously. This complexity implies potentially numerous and diverse mechanisms by which regulatory variants act on gene regulation. The authors set out to find evidence for one of these mechanisms, namely perturbation of TF binding sites. They begin by showing that variation in histone modifications has a strong genetic basis and that the polymorphisms influencing these marks overlap with known regulatory variants such as eQTLs. They then show that polymorphisms in TF binding sites associate with changes in histone marks, providing evidence for directionality in the relationship between these marks and gene regulation. In essence, their results suggest that histone modifications are directed, at least in part, by TF binding. Finally, they find that regulatory variants can have an impact on the molecular phenotypes of distal regions.

I found this paper, as well as the other three previously mentioned, to be quite interesting. I think these papers show that our understanding of gene regulation is still very simplistic. With the advent of high-throughput molecular assays like ChIP-Seq and DNase-Seq, we can begin to interrogate the complex role of regulatory variation on many phenotypes. In doing so, it is of primary interest to ask questions regarding directionality. How do a given set of molecular phenotypes relate? Do these phenotypes represent a cause or a consequence of genome function? How do the diverse elements of gene regulation function together to build complex phenotypes?

References

[1] Kilpinen, H., Waszak, S. M., Gschwind, A. R., Raghav, S. K., Witwicki, R. M., Orioli, A., Dermitzakis, E. T., et al. (2013). Coordinated Effects of Sequence Variation on DNA Binding, Chromatin Structure, and Transcription. Science (New York, N.Y.), 744. doi:10.1126/science.1242463

[2] McVicker, G., van de Geijn, B., Degner, J. F., Cain, C. E., Banovich, N. E., Raj, A., Pritchard, J. K., et al. (2013). Identification of Genetic Variants That Affect Histone Modifications in Human Cells. Science (New York, N.Y.), 747. doi:10.1126/science.1242429

[3] Kasowski, M., Kyriazopoulou-Panagiotopoulou, S., Grubert, F., Zaugg, J. B., Kundaje, A., Liu, Y., Snyder, M., et al. (2013). Extensive Variation in Chromatin States Across Humans. Science (New York, N.Y.), 750. doi:10.1126/science.1242510

[4] Heinz, S., Romanoski, C. E., Benner, C., Allison, K. a, Kaikkonen, M. U., Orozco, L. D., & Glass, C. K. (2013). Effect of natural genetic variation on enhancer selection and function. Nature, 503(7477), 487–492. doi:10.1038/nature12615

[5] Pickrell, J. K., Marioni, J. C., Pai, A. a, Degner, J. F., Engelhardt, B. E., Nkadori, E., Pritchard, J. K., et al. (2010). Understanding mechanisms underlying human gene expression variation with RNA sequencing. Nature, 464(7289), 768–72. doi:10.1038/nature08872

[6] Degner, J. F., Pai, A. a, Pique-Regi, R., Veyrieras, J.-B., Gaffney, D. J., Pickrell, J. K., Pritchard, J. K., et al. (2012). DNase I sensitivity QTLs are a major determinant of human expression variation. Nature, 482(7385), 390–4. doi:10.1038/nature10808

Paper author Jonathan Pritchard is a professor in the Departments of Genetics and Biology.

Paper author Jonathan Pritchard is a professor in the Departments of Genetics and Biology.