Monday 25 November 2013

Cellular Senescence and Genome Wide Methylation Analysis

Recently we conducted a study into the pattern of methylation within senescent cells.

Senescence

Senescence is the permanent halt of proliferation associated with cellular stresses including DNA damage, damaged or short telomeres and oxidative stress.  Senescence dates back to the Hayflick limit and, while there are no definitive markers, cells typically become large and flat and can be identified by the presence of SA-β-GAL, SAHF and tumour suppression networks (i.e. p53 / p16).

Senescence is thought to be a tumour suppression mechanism, through blocking clonal proliferation of cancer cells, and a promoter of ageing via an inflammaging phenotype.  Cancer cells must somehow bypass senescence in order to continue proliferating.


Methylation 

We were interested in studying genome-wide DNA methylation patterns within senescence.  DNA methylation is the addition of a Methyl group to specific DNA bases.  In mammals this is restricted to the cytosine nucleotide (C) within DNA.  Methylation is largely localised at CpG's (a cytosine nucleotide followed by a guamine nucleotide) within the genome.

In order to study DNA methylation we use whole genome sequencing combined with bisulphite treatment.  Bisulphite treatment of DNA converts unmethylated cytosines into uracil - when we apply a process known as polymerase chain reaction (PCR) to the DNA before sequencing this is then further converted into thymine.  Methylated cytosines are unaffected.  Therefore, when sequencing a methylated cytosine will be read as a cytosine while an unmethylated cytosine will be read as a thymine.  As such when we sequence the DNA we can then deduce, via comparison with the human reference genome, which cytosine nucleotides were originally methylated and which were unmethylated.


Loss of Methylation at Late Replicating Regions

We sequenced three replicates of proliferating and replicative senescent cells (IMR90 cell culture). Our first observation was that there was a general loss of methylation (hypo-methylation) within the cells and that this loss was focused at regions of the genome which were only partially methylated (as opposed to completely unmethylated or fully methylated).  Further investigation highlighted a relationship with the regions losing methylation and late replicating regions of the genome.  i.e. regions of the genome where DNA is replicated later on during the cell cycle[1].


Focal Hypermethylation

We observed that methylation does not globally regulate gene expression changes but we did detect a subset of genes with very similar patterns in terms of methylation changes and which were coregulated at the expression level. In these genes we observed focal increase in methylation (hyper-methylation) at the regions adjacent to islands of high CpG content and that these genes were frequently associated with cell cycle functions.


Relationship to Cancer

Surprisingly, considering the role of senescence as a tumour suppression mechanism, we discovered that the methylation changes in senescence were very similar to the methylation changes that had previously been described in cancers[2,3,4].


Origins

We discovered that the origin of the loss of methylation appeared to be due to mislocalisation of the DNMT1 methyltransferase which is responsible for propagating the original methylation pattern of DNA at replication.  Specifically we noticed that foci of the DNMT1 protein are lost in old cycling cells and that these foci had previously been linked to late replicating regions[5].

Bypassing Senescence

Finally, we bypassed senescence via the introduction of SV40 T antigen into the cells in order to deactivate tumour suppression networks.  This allows the cells to continue proliferating, for a period of time, beyond the usual senescence time limit.  When we did this we made the same measurements of DNA methylation as before and found that the methylation changes that occur on the path to senescence are retained in the bypass cells.


Conclusion 

We therefore conclude that the failure of DNMT1 to methylate late replicating regions in near senescent cells may be the origin of the same methylation patterns found in cancer cells.  Furthermore, given the similarities, it may be possible that these changes may lead cells to be more susceptible to transformation into cancer cells.




The full paper is available online from Nature Cell Biology, is discussed by a CRUK press release and the raw data is available from GEO GSE48580.

References


[1] R. S. Hansen et al., Sequencing newly replicated DNA reveals widespread plasticity in human replication timing, PNAS, 2010.

[2] K. D. Hansen et al., Increased methylation variation in epigenetic domains across cancer types, Nature Genetics, 2011.

[3] B. P. Berman et al., Regions of focal DNA hypermethylation and long-range hypomethylation in colorectal cancer coincide with nuclear lamina–associated domains, Nature Genetics, 2012.

[4] G. C. Hon et al., Global DNA hypomethylation coupled to repressive chromatin domain formation and gene silencing in breast cancer. Genome Research, 2012.

[5] L. Schermelleh  et al., Dynamics of Dnmt1 interaction with the replication machinery and its role in postreplicative maintenance of DNA methylation. Nucleic Acids Res, 2007.

Friday 1 November 2013

What is a replicate anyway?

In science we perform replication as the primary way of ensuring reproducibility. This is one of the key principles. Science only advances through consensus - initially within a single lab and later within the scientific community at large.

However,  there are lots of different ways the replicates can be performed in any given experiment and the choice depends on exactly what experimental conditions you are trying to control for.

In this post I will argue there is a hierarchy of replication which should be considered when planning any scientific experiment.  For the purposes of this post I will use experiments within the molecular biology domain to illustrate my points but I believe it is applicable into other fields.

Technical replicates

Technical replicates are aimed at solving the problem of inaccurate measuring devices.  No measuring device is perfect and they will produce slightly different answers each time you use them.

A typical example of technical replication might be to measure the RNA abundance of a particular gene three times on a single sample.  Subsequently this allows you to make a higher confidence estimate of the actual abundance of the sample - specifically it allows you to actually estimate the degree of error in the measurement. i.e. How accurate is your gene expression abundance measurement anyway?

Experimental replicates

Experimental replicates are used to detect the degree of variation that is introduced by a particular instance of an experiment.  If any given experiment is repeated on a different day it will have slightly different results from the original.  Experimental replicates are there to figure out exactly how big an effect this has on your experimental observations.  Experimental replicates may well be performed on the same tissue sample or cell line if the aim is to show the effect of a particular treatment is consistent for a sample.

Examples of things that might influence how well an experiment replicates on different days might include environmental effects such as ambient temperature, the experimenters mood (and subsequently how careful they were) or how many interruptions they had and when they were.

Biological replicates

Next up is biological replication which aims to show that the experimental observation is generalisable to more than a specific tissue sample or cell line.  For this we need to show that the effect holds across multiple different samples.

The more different the samples are the stronger the case that can be made.  For example if an effect is shown to be consistent across a population of genetically identical mice this is not as strong as showing that the effect is the same in another different strain of mouse.  Even stronger would be using different rodents or even a completely different species altogether.

Experimenter replicates

Each experimenter introduces their own variation via differences in skill, patience and familiarity with a given assay.  Replication by multiple people shows that the result can be obtained by a different person using the same protocol, reagents and equipment as the original experiment.

For complicated or tricky experiments it is often common for a researcher to visit a lab which has the experiment working in order to learn first-hand the techniques used by the original experimenter to enable them to be able to reproduce an experiment on their own.

Method replicates

Finally, another approach to confirming your results is to use a different method entirely.  In most high impact papers it is often the case that an important observation may be shown in multiple different ways and in multiple model systems in order to demonstrate its robustness.

One simple example might be to show the change in regulation of a gene or pathway via detection using a whole genome screen (RNA-seq/microarray), quantitative RNA abundance (qRT-PCR), protein abundance (western blot) and via changes in localisation within the tissue (immunohistochemistry).  These methods may therefore be complimentary to one another and tell the same story but share little otherwise in common.  This may be extended to show the effect in a variety of tissues or model systems.

Laboratory replicates

Laboratory replicates are the gold standard of experimental replication and are used to demonstrate that the results of an experiment can be replicated by the scientific community at large.  When results are replicated by many different laboratories they come as close to the label of "truth" that science can reasonably apply.

Variation between labs and institutes can be induced by apparently minor differences in protocol, equipment, reagents, regulations, behaviour, strains of model system, batches of the same reagent and suppliers of equipment.

Summary

As you move up the replication hierarchy, the number of confounding variables will increase - your experiment will be less controlled.  But the results will be more generalisable and robust.  Ultimately the strength of your observation should exceed the noise introduced by any of the confounding variables we've discussed above in order to be accepted by the scientific community. Otherwise it just goes down in history as unreproducible.

But how easy should it be to reproduce results?  Methods sections may be incomplete and this not necessarily the authors fault; journals frequently impose word limits which may preclude full detail.  Reagents used may no longer be available and the replacement may not work in the same way.  And there is just as much chance that the original experiment is correct and the replication is wrong as there is of the reverse.  For a more in depth discussion see this post by

It is good to see endeavours such as the reproducibility initiative set out to independently reproduce important papers.  In addition important papers will be reproduced simply because other labs will want to build on previous published results and will discover problems in reproducing it while they try to extend it.

This does, however, leave one problem.  Small papers, with less ground-breaking results, published in low-tier journals are less likely to be reproduced by other labs - they may not be as interesting for other labs to build on. As such their results may live on in science as a dogma that may persist for many years.  It is unclear to me how to address this final issue as there are currently very few incentives for scientists to reproduce results for their own sake.

Tuesday 16 July 2013

The G Test and Bisulphite Sequencing DMRs

This is Part 4 in a series on Bisulphite Sequencing.
You can return to Part 3 (Bisulphite Sequencing Simple Visualisation).


Okay, so we have our bisulphite sequencing data on the UCSC genome browser. Lets assume we have two conditions CondA and CondB, and we have two replicates of each condition. We can see what appear to be changes from our visual inspection on the browser. Now what?

Well, we probably want to compute differentially methylated regions. With 1 replicate of each that's fairly straight forward for categorical data like the bisulphite seq data we have - we use fishers exact test. With replicates a simple approach is just to pool the two replicates together and do our fishers exact test on the pooled data, but this doesn't give us a very good idea of what's going on if the two replicates disagree for a given region.

Instead we use the G-test which can be calculated for any given contingency table with the following formula:

$$$$ G = 2 \sum_{i=0}^{r} \sum_{j=0}^{c} {O_{i,j} \ln(O_{i,j}/E_{i,j}) } $$$$
Where, $$$r$$$ and $$$c$$$ are the number of rows and columns in the contingency table, $$$O_{i,j}$$$ and $$$E_{i,j}$$$ are the Observed and Expected values for cell $$$i,j$$$ in the table. I've detailed how to implement the G-test for a 2x2 contingency table in Python previously.

The main selling point of the G-test, over a regular chi-square test, is that the G value is additive. Consider the following:

G-test 1G-test 2
A B G-value
C 100 70 = 7.1
D 35 50
A B G-value
C 200 140 = 14.2
D 70 100


To find out where we have differential regions we first partition the genome into 2kb windows. There are large regions of the genome where there are no CpG's or, alternatively, where we have no sequencing coverage - we will skip those regions. We align our windows such that each window starts at an individual CpG and stops at the last CpG within 2kb of the start position.

Pooled G

We first compute a Pooled G which is the G-test performed on the pooled replicates. The contingency table would consist of columns representing each condition and rows representing methylated and unmethylated counts. The number of degrees of freedom is equal to $$$(R-1) \cdot (C-1)$$$ so as we are doing a single 2x2 contingency table this will always be equal to 1.

def pooledG(lmeth,lunmeth,rmeth,runmeth):   
    dof = 1
    pooledg = gtest(lmeth,lunmeth,rmeth,runmeth)
    return pooledg, dof, GtoP(pooledg,dof)


Total G

We then compute a G-value for each individual replicate and add them together. Each G-test we do adds 1 to the total number of degrees of freedom.

def totalG(data):
    dof = 0
    totalg = 0
    for lmeth,lunmeth,rmeth,runmeth in data:
            totalg += gtest(lmeth,lunmeth,rmeth,runmeth)
            dof+=1 
    return totalg, dof, GtoP(totalg,dof)


Heterogeneous G

If each of our three replicates reproduce exactly then Pooled G will be identical to Total G (due to the additive nature of G-test results). However, if the ratios are not identical then they will not match. The difference between the Total G-value and the Pooled G-value is the heterogeneous G-value and tells us if the difference between our replicates is significant (i.e. our replicates are not all doing the same thing.

def heterogenousG(totalg,totaldof,pooledg,pooleddof):
    heteroG = totalg-pooledg
    heterodof = totaldof-pooleddof
    return heteroG,heterodof,GtoP(heteroG,totaldof-pooleddof)


Putting it all together

Now, we can iterate through a file summarising the number of methylated and unmethylated reads at each CpG (described in a previous post) and compute the Pooled, Total and Heterogenous G-values (and corresponding p-values) for each genomic window.
import getopt
import sys
import csv
from itertools import izip_longest
from itertools import izip
from memoized import memoized

# sudo easy_install mpmath
import mpmath
#mpmath.mp.dps = 20 # decimal digits of precision

@memoized
def mpmathcdf(g,df,dps=10):
    mpmath.mp.dps = dps
    
    x,k = mpmath.mpf(g), mpmath.mpf(df)
    cdf = mpmath.gammainc(k/2, 0, x/2, regularized=True)
    
    # floating point precision insufficient, use more precision
    if cdf == 1.0:
        if dps > 4000:
            return cdf # give up after a certain point
        else:
            cdf = mpmathcdf(g,df,dps*2)
    return cdf

def GtoP(g,df):
    assert g >= 0, g
    return float(1-mpmathcdf(g,df))

import math

@memoized
def flnf(f):
    return f*math.log(f) if f>0.5 else 0

@memoized
def gtest(a,b,c,d):
    row1 = a+b
    row2 = c+d
    col1 = a+c
    col2 = b+d
    
    total = flnf(a+b+c+d)
    celltotals = flnf(a)+flnf(b)+flnf(c)+flnf(d)
    rowtotals = flnf(row1)+flnf(row2)
    coltotals = flnf(col1)+flnf(col2)
    
    # abs is necessary to account for float precision errors when doing the subtraction
    return abs(2*((celltotals + total)-(rowtotals+coltotals)))



# extract the pooled meth counts from the sample
def pooledMethCounts(data,covs,compare):
    lmethtotal = 0
    lunmethtotal = 0
    rmethtotal = 0
    runmethtotal = 0
    
    if len(compare)==2:
        # if we are given covariates to use
        covleft, covright = compare    
        for cov, (meth,unmeth) in izip(covs,grouper(2,data)): # extract in pairs of meth/unmeth
            if cov == covleft:
                lmethtotal += meth
                lunmethtotal += unmeth
            elif cov == covright:
                rmethtotal += meth
                runmethtotal += unmeth
    else:
        # if we have not been given them, assume that each adjacent sample (which is one pair) is one replicate
        for lmeth,lunmeth,rmeth,runmeth in grouper(4,data): # extract in quads (2 pairs) representing one replicate
            lmethtotal += lmeth
            lunmethtotal += lunmeth
            rmethtotal += rmeth
            runmethtotal += runmeth
        
    return lmethtotal,lunmethtotal,rmethtotal,runmethtotal


# extract individual replicate counts from the sample and return as a list of (lmeth,lunmeth,rmeth,runmeth)
# where l and r refer to the left and right cov in the compare array
# if compare array is empty assume that data is formatted as:
# lmeth1,lunmeth1,rmeth1,runmeth1,lmeth2,lunmeth2,rmeth2,runmeth2 etc
def covReplicateMethCounts(data,covs,compare):
    if len(compare)==2:
        covleft, covright = compare
        
        lreplicates = []
        rreplicates = []
    
        for cov, (meth,unmeth) in izip(covs,grouper(2,data)): # extract in pairs of meth/unmeth
            if cov == covleft:
                lreplicates.append((meth,unmeth))
            elif cov == covright:
                rreplicates.append((meth,unmeth))

        #return [(lmeth,lunmeth,rmeth,runmeth),(),(),...]    
        return [(a,b,c,d) for (a,b),(c,d) in izip(lreplicates,rreplicates)]
    else:
        # assume that each pair of samples (i.e. 4 columns) in the input is one replicate
        # in this case our input file is in the right order and we just need to split it up.
        return grouper(4,data)


# total G: at least some of the data is different
# 1 degree of freedom per replicate (i.e. 3 in total), add G's from each seperate test
def totalG(data,covs,compare):
    dof = 0
    totalg = 0
    for lmeth,lunmeth,rmeth,runmeth in covReplicateMethCounts(data,covs,compare): # extract in quads (2 pairs)
            totalg += gtest(lmeth,lunmeth,rmeth,runmeth)
            dof+=1 
    assert totalg >= 0
    return totalg, dof, GtoP(totalg,dof)

# pooled G: the data as a whole is different
# 1 degree of freedom, pool results from all replicates and do 1 2x2 table
def pooledG(data,covs,compare):    
    lmeth,lunmeth,rmeth,runmeth = pooledMethCounts(data,covs,compare)
    
    dof = 1
    
    pooledg = gtest(lmeth,lunmeth,rmeth,runmeth)
    
    assert pooledg >= 0
    
    return pooledg, dof, GtoP(pooledg,dof)

# there is evidence that they are significantly different from each other
# heterogeneity G: total G - pooled G, df = 3-1 = 2
def heterogeneityG(totalg,totaldof,pooledg,pooleddof):
    heteroG = max(totalg-pooledg,0)
    heterodof = totaldof-pooleddof
    return heteroG,heterodof,GtoP(heteroG,totaldof-pooleddof)


def grouper(n, iterable, fillvalue=None):
    "grouper(3, 'ABCDEFG', 'x') --> ABC DEF Gxx"
    args = [iter(iterable)] * n
    return izip_longest(fillvalue=fillvalue, *args)

def unpackLine(chrm,coord,*data):
    return chrm.strip(), int(coord), [int(d) for d in data]

def extractLine(line):
    chrm,coord,data = unpackLine(*line)
    if not chrm.startswith("chr"):
        chrm = "chr"+chrm
    return (chrm,coord,data)


    
if __name__ == '__main__':
    try:
        opts, args = getopt.getopt(sys.argv[1:], "", ["combined=","cov=","compare="])
    except getopt.GetoptError, err:
        # print help information and exit:
        print str(err) # will print something like "option -a not recognized"
        sys.exit(2)
    
    infile = None
    covs = None
    compare = None
    
    for o, a in opts:
        if o=="--combined":
            infile = a
            print "combined", a
        elif o=="--cov":
            covs = a.split(",")
            print covs
        elif o=="--compare":
            compare = a.split(",")
            print compare
            assert len(compare)==2
    
    assert infile != None
    
    # we either have compare and covs or neither
    assert compare == None or len(compare)==2
    if compare == None: assert covs == None
    if len(compare)==2: assert covs != None
    
    methfile = csv.reader(open(infile, "r"), delimiter="\t")
    
    # ---    
    # windows
    
    if compare != None:
        leftcov,rightcov = compare
        windows = csv.writer(open(infile+"."+leftcov+"-vs-"+rightcov+".windows.bedGraph","w"),delimiter='\t')
        percent = csv.writer(open(infile+"."+leftcov+"-vs-"+rightcov+".percent.windows.bedGraph","w"),delimiter='\t')
    else:
        windows = csv.writer(open(infile+".windows.bedGraph","w"),delimiter='\t')
        percent = csv.writer(open(infile+".percent.windows.bedGraph","w"),delimiter='\t')
    
    windowSize = 2000
    
    currentWindowPos = None
    currentChr = None
    
    cs = 0
    dataCount = None
    
    def writeWindow(chrm,pos,epos):
        global dataCount
        global cs
        global windows
        global percent
        
        lmethCount,lunmethCount,rmethCount,runmethCount = pooledMethCounts(dataCount,covs,compare)
        
        if (lmethCount + lunmethCount) == 0 or (rmethCount + runmethCount) == 0:
            print "Unprintable window:",chrm, pos, epos, lmethCount, lunmethCount, rmethCount, runmethCount
            return
        
        lPercentage = (float(lmethCount)/float(lmethCount+lunmethCount)) * 100.0
        rPercentage = (float(rmethCount)/float(rmethCount+runmethCount)) * 100.0
        
        total,totaldof,totalp = totalG(dataCount,covs,compare)
        pooled,pooleddof,pooledp = pooledG(dataCount,covs,compare)
        hetero,heterodof,heterop = heterogeneityG(total,totaldof,pooled,pooleddof)
        
        windows.writerow([chrm,pos,epos+1,'1' if rPercentage > lPercentage else '-1',totalp,pooledp,heterop,cs])
        
        percent.writerow([chrm,pos,epos+1,lPercentage,rPercentage])         
        
        dataCount = None
        cs = 0

    lastCoordCounted = None
    # the actual work here
    for line in methfile:
        
        chrm,coord,data = extractLine(line)

        if currentChr == None:
            currentChr = chrm
            currentWindowPos = coord
            lastCoordCounted = coord
        else:
            # always write a window if we change chrm
            if chrm != currentChr:
                writeWindow(currentChr,currentWindowPos,lastCoordCounted)
                currentChr = chrm
                currentWindowPos = coord
                
            # write a window if the next cpg is outside the current window
            elif coord - currentWindowPos > windowSize:
                writeWindow(currentChr,currentWindowPos,lastCoordCounted)
                currentChr = chrm
                currentWindowPos = coord
                
            # otherwise dont write any window (the current cpg is in the window)
        
        ####
        # add data to existing data count
        ####
        if dataCount == None:
            dataCount = data
        else:
            for index,d in enumerate(data):
                dataCount[index] += d
        lastCoordCounted = coord
        
        # inc number of c's counted
        cs += 1
        
    # after we write a window there is always at least one unmeasured coord after it so we have one more window to write
    writeWindow(currentChr,currentWindowPos,lastCoordCounted)


Note that the above script relies upon the Memoized decorator.

Executed as:

python replicatesGTest.py \
--combined "Conditions.1.and.2.CpG_context.aggregate" \
--cov Condition1,Condition2,Condition1,Condition2,Condition1,Condition2
--compare Condition1,Condition2 


Where --cov lists the order of the conditions in the file and --cov indicates which conditions should be compared.

For each window, this script outputs a tab separated file containing the chromosome,start and stop position of the window,'1' if it is "increasing" otherwise '-1' and p-values for computed from the Total, Pooled ahd Hetero G computations as well as the number of methylation sites considered in the window.

Note that the p-values generated are uncorrected for multiple testing and should be FDR corrected. I've previously described a suitable approach for large files such as these.

Differentially methylated regions can then be defined as regions of adjacent windows with a FDR adjusted p-value less than or equal to a fixed level (i.e. 0.05 FDR) that change in the same direction.

Tuesday 9 July 2013

The G-Test and Python

The G-Test is a statistical test which can be used much like a chi-squared test. It is defined as: $$$$ G = 2 \sum_{i=0}^{r} \sum_{j=0}^{c} {O_{i,j} \ln(O_{i,j}/E_{i,j}) } $$$$ Where, $$$r$$$ and $$$c$$$ are the number of rows and columns in the contingency table and $$$O_{i,j}$$$ and $$$E_{i,j}$$$ are the Observed and Expected values for cell $$$i,j$$$ in the table.

To make a straightforward implementation for it in Python we first do a bit of math. For a 2x2 contingency table:

$$$ G = 2 \Bigg({O_{1,1} \ln(\frac{O_{1,1}}{E_{1,1}})} + {O_{1,2} \ln(\frac{O_{1,2}}{E_{1,2}})} + {O_{2,1} \ln(\frac{O_{2,1}}{E_{2,1}})} + {O_{2,2} \ln(\frac{O_{2,2}}{E_{2,2}})} \Bigg) $$$

Expand via log equivalences:

$$$ G = 2 \Bigg( $$$ $$$ \bigg( {O_{1,1} \ln(O_{1,1}) - O_{1,1} \ln(E_{1,1})} \bigg) + \bigg( {O_{1,2} \ln(O_{1,2}) - O_{1,2} \ln(E_{1,2})} \bigg) $$$
$$$ + \bigg( {O_{2,1} \ln(O_{2,1}) - O_{2,1} \ln(E_{2,1})} \bigg) + \bigg( {O_{2,2} \ln(O_{2,2}) - O_{2,2} \ln(E_{2,2})} \bigg) \Bigg) $$$


Group Observed vs Expected together:

$$$ G = 2 \Bigg( $$$ $$$ \bigg( O_{1,1} \ln(O_{1,1}) + O_{1,2} \ln(O_{1,2}) + O_{2,1} \ln(O_{2,1}) + O_{2,2} \ln(O_{1,2}) \bigg)$$$
$$$ - \bigg( O_{1,1} \ln(E_{1,1}) + O_{1,2} \ln(E_{1,2}) + O_{2,1} \ln(E_{2,1}) + O_{2,2} \ln(E_{2,2}) \Bigg) $$$


For a 2x2 contingency table; $$$E_{i,j} = \frac{T_{i} \cdot T_{j}}{T} $$$, where $$$T_{i}$$$ and $$$T_{j}$$$ are the totals for the row and column of cell $$$E_{i,j}$$$ and $$$T$$$ is the total for all cells in the table. Specifically, $$$$ \ln(E_{1,1}) = \ln \Bigg( \frac{ (O_{1,1}+O_{1,2}) \cdot (O_{1,1}+O_{2,1}) }{ O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} } \Bigg) $$$$ $$$$ \ln(E_{1,2}) = \ln \Bigg( \frac{ (O_{1,1}+O_{1,2}) \cdot (O_{1,2}+O_{2,2}) }{ O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} } \Bigg) $$$$ $$$$ \ln(E_{2,1}) = \ln \Bigg( \frac{ (O_{2,1}+O_{2,2}) \cdot (O_{1,1}+O_{2,1}) }{ O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} } \Bigg) $$$$ $$$$ \ln(E_{2,2}) = \ln \Bigg( \frac{ (O_{2,1}+O_{2,2}) \cdot (O_{1,2}+O_{2,2}) }{ O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} } \Bigg) $$$$
We then substitute these into our equation for G.
$$$ G = 2 \Bigg( $$$ $$$\bigg( O_{1,1} \ln(O_{1,1}) + O_{1,2} \ln(O_{1,2}) + O_{2,1} \ln(O_{2,1}) + O_{2,2} \ln(O_{1,2}) \bigg) - $$$
$$$ \bigg( O_{1,1} \ln \Big( \frac{ (O_{1,1}+O_{1,2}) \cdot (O_{1,1}+O_{2,1}) }{ O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} } \Big) + O_{1,2} \ln \Big( \frac{ (O_{1,1}+O_{1,2}) \cdot (O_{1,2}+O_{2,2}) }{ O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} } \Big) + $$$
$$$ O_{2,1} \ln \Big( \frac{ (O_{2,1}+O_{2,2}) \cdot (O_{1,1}+O_{2,1}) }{ O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} } \Big) + O_{2,2} \ln \Big( \frac{ (O_{2,1}+O_{2,2}) \cdot (O_{1,2}+O_{2,2}) }{ O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} } \Big) \bigg) \Bigg) $$$

Another log equivalence gives us this:
$$$ G = 2 \Bigg( $$$ $$$\bigg( O_{1,1} \ln(O_{1,1}) + O_{1,2} \ln(O_{1,2}) + O_{2,1} \ln(O_{2,1}) + O_{2,2} \ln(O_{1,2}) \bigg) - $$$
$$$ \bigg( O_{1,1} \ln ( (O_{1,1}+O_{1,2}) \cdot (O_{1,1}+O_{2,1}) ) - O_{1,1} \ln ( O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} ) + $$$
$$$ O_{1,2} \ln ( (O_{1,1}+O_{1,2}) \cdot (O_{1,2}+O_{2,2}) ) - O_{1,2} \ln ( O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} ) + $$$
$$$ O_{2,1} \ln ( (O_{2,1}+O_{2,2}) \cdot (O_{1,1}+O_{2,1}) ) - O_{2,1} \ln ( O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} ) + $$$
$$$ O_{2,2} \ln ( (O_{2,1}+O_{2,2}) \cdot (O_{1,2}+O_{2,2}) ) - O_{2,2} \ln ( O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} ) \bigg) \Bigg) $$$

Group common factor of $$$\ln ( O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2} )$$$:
$$$ G = 2 \Bigg( $$$ $$$ \bigg( O_{1,1} \ln(O_{1,1}) + O_{1,2} \ln(O_{1,2}) + O_{2,1} \ln(O_{2,1}) + O_{2,2} \ln(O_{1,2}) \bigg) +$$$
$$$ \bigg( (O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2}) \cdot \ln(O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2}) \bigg) - $$$
$$$ \bigg( O_{1,1} \ln \big( (O_{1,1}+O_{1,2}) \cdot (O_{1,1}+O_{2,1}) \big) + O_{1,2} \ln \big( (O_{1,1}+O_{1,2}) \cdot (O_{1,2}+O_{2,2}) \big) + $$$
$$$ O_{2,1} \ln \big( (O_{2,1}+O_{2,2}) \cdot (O_{1,1}+O_{2,1}) \big) + O_{2,2} \ln \big( (O_{2,1}+O_{2,2}) \cdot (O_{1,2}+O_{2,2}) \big) \bigg) \Bigg) $$$

More log equivalences:
$$$ G = 2 \Bigg( $$$ $$$ \bigg( O_{1,1} \ln(O_{1,1}) + O_{1,2} \ln(O_{1,2}) + O_{2,1} \ln(O_{2,1}) + O_{2,2} \ln(O_{1,2}) \bigg) + $$$
$$$ \bigg( (O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2}) \cdot \ln(O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2}) \bigg) -$$$
$$$ \bigg( O_{1,1} \ln (O_{1,1}+O_{1,2}) + O_{1,1} \ln (O_{1,1}+O_{2,1}) + O_{1,2} \ln (O_{1,1}+O_{1,2}) + O_{1,2} \ln (O_{1,2}+O_{2,2}) + $$$
$$$ O_{2,1} \ln (O_{2,1}+O_{2,2}) + O_{2,1} \ln (O_{1,1}+O_{2,1}) + O_{2,2} \ln (O_{2,1}+O_{2,2}) + O_{2,2} \ln (O_{1,2}+O_{2,2}) \bigg) \Bigg) $$$

Group common factors of $$$\ln (O_{1,1}+O_{1,2})$$$, $$$\ln (O_{1,1}+O_{2,1})$$$ etc. Yielding:
$$$ G = 2 \Bigg( $$$ $$$ \bigg( O_{1,1} \ln(O_{1,1}) + O_{1,2} \ln(O_{1,2}) + O_{2,1} \ln(O_{2,1}) + O_{2,2} \ln(O_{1,2}) \bigg) $$$
$$$ + \bigg( (O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2}) \cdot \ln(O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2}) \bigg) $$$
$$$ - \bigg( (O_{1,1} + O_{1,2}) \ln (O_{1,1}+O_{1,2}) + (O_{2,1} + O_{2,2}) \ln (O_{2,1}+O_{2,2}) \bigg) $$$
$$$ - \bigg( (O_{1,1} + O_{2,1}) \ln (O_{1,1}+O_{2,1}) + (O_{1,2} + O_{2,2}) \ln (O_{1,2}+O_{2,2}) \bigg) \Bigg) $$$

We can then summarise the four parts of the above equation as follows:
$$$$ celltotals = O_{1,1} \ln(O_{1,1}) + O_{1,2} \ln(O_{1,2}) + O_{2,1} \ln(O_{2,1}) + O_{2,2} \ln(O_{1,2}) $$$$ $$$$ total = (O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2}) \ln(O_{1,1} + O_{1,2} + O_{2,1} + O_{2,2}) $$$$ $$$$ rowtotals = (O_{1,1} + O_{1,2}) \ln (O_{1,1}+O_{1,2}) + (O_{2,1} + O_{2,2}) \ln (O_{2,1}+O_{2,2}) $$$$ $$$$ columntotals = (O_{1,1} + O_{2,1}) \ln (O_{1,1}+O_{2,1}) + (O_{1,2} + O_{2,2}) \ln (O_{1,2}+O_{2,2}) $$$$ where: $$$$ G = 2 \big(celltotals + total - (rowtotals + columntotals) \big) $$$$ If we assign $$$a = O_{1,1}, b = O_{1,2}, c = O_{2,1}, d = O_{2,2}$$$ then we can then implement as follows in Python:

import math

def flnf(f):
    return f*math.log(f) if f>0.5 else 0

def gtest(a,b,c,d):
    row1 = a+b
    row2 = c+d
    col1 = a+c
    col2 = b+d
    
    total = flnf(a+b+c+d)
    celltotals = flnf(a)+flnf(b)+flnf(c)+flnf(d)
    rowtotals = flnf(row1)+flnf(row2)
    coltotals = flnf(col1)+flnf(col2)
    
    return 2*(celltotals + total-(rowtotals+coltotals))


We then probably want to convert our G-value into a P-value. The G-test approximates a chi squared distribution and has the same number of degrees of freedom.

If the calculated chi-squared distribution cumulative distribution function is equal to exactly 1 (i.e. a p-value of exactly 0) then we increase the precision and try again. This means that we can find the correct order of magnitude for our p-value (which is good for when we FDR correct these p-values later on) while minimising the number of calculations we need to do (the majority of conversions don't need 4000 decimal places computed). ($$$g$$$ is the G-value and $$$df$$$ is the degrees of freedom - 1 for a 2x2 table).

import mpmath

def mpmathcdf(g,df,dps=10):
    mpmath.mp.dps = dps
    
    x,k = mpmath.mpf(g), mpmath.mpf(df)
    cdf = mpmath.gammainc(k/2, 0, x/2, regularized=True)
    
    # floating point precision insufficient, use more precision
    if cdf == 1.0:
        if dps > 4000:
            return cdf # give up after a certain point
        else:
            cdf = mpmathcdf(g,df,dps*2)
    return cdf

def GtoP(g,df):
    assert g >= 0, g
    return float(1-mpmathcdf(g,df))


Update:

Following on from the comments below I've written up a generalised version of the G-test for any sized contingency table which uses the trick from above. I've not proven explicitly that this is necessarily generalisable for any sized table but I'm pretty sure it is.
def gtest(a):
    rowtotals = []
    coltotals = [0]*len(a[0])
    cells = []
    for row in a:
        rowtotals.append(sum(row))
        for index,col in enumerate(row):
            coltotals[index] += col
            cells.append(col)
    return 2*(math.fsum([flnf(x) for x in cells]) + flnf(sum(cells)) - ( math.fsum([flnf(x) for x in rowtotals]) + math.fsum([flnf(x) for x in coltotals]) ))

For a table with 2 columns and 3 rows this can then be called as follows:
gtest([[268,807],
       [199,759],
       [42,184]])

> 7.300817075862142

Again, because I've not proven this will always work it should, of course, be used at your own risk. That said, the same technique is used for up to 10 columns and 50 rows in the G-test spreadsheet at the Handbook for Biological Statistics so it's probably safe at least that far.

Monday 8 July 2013

Bisulphite Sequencing Simple Visualisation (Multiple Samples)

This is Part 3 in a series on Bisulphite Sequencing.
You can return to Part 2 (Using the Binomial Distribution in Bisulphite Sequencing).
or skip to Part 4 (The G Test and Bisulphite Sequencing DMRs).


Jumping back to bisulphite sequencing analysis for a bit.  After we have summarised our data by individual CpG we probably want to do something useful with it.

First up, lets just display it on the UCSC genome browser.  When we display bisulphite sequening data on UCSC we probably want something like this below:


We also want to filter out low confidence CpGs (those with less than a certain number of reads). bismarkToBedGraph does something pretty much like this.

However, it is important that when we filter a particular CpG from any given sample we should also filter it from the other samples displayed at the same time - otherwise we give a misleading indication that there is 0% methylation in the filtered sample and a non-negative amount of methylation in the unfiltered sample.

To address this I made a short script that can take multiple replicates/samples of bisulphite sequencing data, summarised by individual CpGs, and output bedGraph files such that the same CpGs are always reported based on coverage.

import getopt
import sys
import csv
from itertools import izip_longest

class CommentException(Exception):
    pass

def grouper(n, iterable, fillvalue=None):
    "grouper(3, 'ABCDEFG', 'x') --> ABC DEF Gxx"
    args = [iter(iterable)] * n
    return izip_longest(fillvalue=fillvalue, *args)

def unpackLine(chrm,coord,*data):
    return chrm.strip(), int(coord), [int(d) for d in data]

def extractLine(line):
    if line[0].startswith("#"):
        raise CommentException()
    chrm,coord,data = unpackLine(*line)

    if not chrm.startswith("chr"):
        chrm = "chr"+chrm
    return (chrm,coord,data)
    
if __name__ == '__main__':
    try:
        opts, args = getopt.getopt(sys.argv[1:], "", ["combined=","suffixes="])
    except getopt.GetoptError, err:
        # print help information and exit:
        print str(err) # will print something like "option -a not recognized"
        sys.exit(2)
    
    infile = None
    suffixes = None
    rightsuffix = None
    
    for o, a in opts:
        if o=="--combined":
            infile = a
            print "combined", a
        elif o=="--suffixes":
            suffixes = a.split(",")
            print "Suffix", suffixes

    
    assert infile != None
    assert suffixes != None
    
    methfile = csv.reader(open(infile, "r"), delimiter="\t")
    
    files = {}
    
    for suffix in suffixes:
        percentage = csv.writer(open(infile+"-"+suffix+".percent.bedGraph","w"),delimiter='\t')
        coverage = csv.writer(open(infile+"-"+suffix+".coverage.bedGraph","w"),delimiter='\t')
        files[suffix]=(percentage,coverage)
    
    # ---
    # for each
    # coverage
    # percentage
    
    minReadsForPercentage = 5
    minSamplesWithCoverage = 1

    for line in methfile:
        try:
            (chrm,coord,data) = extractLine(line)
        except CommentException:
            continue
        
        percentages = []
        coverages = []
        
        for meth,unmeth in grouper(2,data):
            coverage = meth + unmeth
            if coverage == 0:
                percentage = 0
            else:           
                percentage = (float(meth)/float(meth+unmeth)) * 100.0
            percentages.append(percentage)
            coverages.append(coverage)
    
        satisfycoveragecriteria = 0
        for coverage in coverages:
            if coverage >= minReadsForPercentage:
                satisfycoveragecriteria += 1
        if satisfycoveragecriteria < minSamplesWithCoverage:
            continue

        for index,(percentage,coverage) in enumerate(izip_longest(percentages,coverages)):
            
            percentfile,coveragefile = files[suffixes[index]]
            percentfile.writerow([chrm,coord,coord+1,percentage])
            coveragefile.writerow([chrm,coord,coord+1,coverage])
            

Executed as:

python bisulphiteToBedGraph.py \
--combined "Samples.1.and.2.CpG_context.aggregate" \
--suffixes Sample1,Sample2,Sample3,Sample4 

You can then use bedGraphToBigWig to create bigWig files suitable for UCSC custom track / track hubs.

Wednesday 19 June 2013

qPCR normalisation

This post will discuss the normalisation of qPCR results using both the Delta Delta CT (Livak) and Standard Curve (Pfaffl) methods as applied to qPCR of RNA and ChIP.  It will also discuss some of the common pitfalls when considering replicates of qPCR data.

While this isn't a cutting edge technique, the motivation for this post is the number of qPCR spreadsheets I see that do this wrongly.

Delta Delta CT


The Livak method is more commonly known as the "Delta Delta CT" (ΔΔCT). The Delta Delta CT method makes one important assumption about the PCR, namely, the amplification efficiencies of the reference control gene and the target gene of interest must be approximately equal.  Specifically Delta Delta CT assumes that each PCR cycle will exactly double the amount of material in your sample (amplification efficiency = 100%).

$$$ ΔΔCT = ΔCT (treated sample) - ΔCT (untreated sample) $$$

where $$$ ΔCT(sample) = CT(target) - CT(ref) $$$, therefore

$$$ ΔΔCT = (CT(target,untreated) - CT(ref,untreated)) - (CT(target,treated) - CT(ref,treated)) $$$

where

$$$CT(target,untreated)$$$ = CT value of gene of interest in untreated sample
$$$CT(ref,untreated)$$$ = CT value of control gene in untreated sample
$$$CT(target,treated)$$$ = CT value of gene of interest in treated sample
$$$CT(ref,treated)$$$ =  CT value of control gene in treated sample

We can then calculate the ratio of our target gene in our treated sample relative to our untreated sample by taking $$$2^{ΔΔCT}$$$.

A quick worked example:

UntreatedTreated
Ref Gene16.1715.895
Target Gene21.22519.763


$$$$ΔΔCT = (CT(target,untreated) - CT(ref,untreated)) - (CT(target,treated) - CT(ref,treated))$$$$
$$$$ΔΔCT = (21.225 - 16.17) - (19.763 - 15.895)$$$$
$$$$ΔΔCT = (5.055) - (3.868)$$$$
$$$$ΔΔCT = 1.187$$$$
$$$$2^{ΔΔCT} = 2^{1.187} = 2.277$$$$

So our gene of interest is increased by 2.277 times in our treated sample versus our untreated sample.

The exact order that you do the subtraction in actually doesn't make a huge difference and you'll probably see other people do it slightly differently.  Consider these possibilities:

ΔUntreated vs ΔTreated

$$$ΔΔCT = (CT(target,untreated) - CT(ref,untreated)) - (CT(target,treated) - CT(ref,treated))$$$
$$$ΔΔCT = (21.225 - 16.17) - (19.763 - 15.895)$$$
$$$ΔΔCT = 1.187$$$

$$$ΔΔCT = (CT(ref,treated) - CT(target,treated)) - (CT(ref,untreated) - CT(target,untreated))$$$
$$$ΔΔCT = (15.895 - 19.763) - (16.17 - 21.225)$$$
$$$ΔΔCT = 1.187$$$

$$$ΔΔCT = (CT(ref,untreated) - CT(target,untreated)) - (CT(ref,treated) - CT(target,treated))$$$
$$$ΔΔCT = (16.17 - 21.225) - (15.895 - 19.763)$$$
$$$ΔΔCT = -1.187$$$

$$$ΔΔCT = (CT(target,treated) - CT(ref,treated)) - (CT(target,untreated) - CT(ref,untreated))$$$
$$$ΔΔCT = (19.763 - 15.895) - (21.225 - 16.17)$$$
$$$ΔΔCT = -1.187$$$


ΔReference Gene vs ΔTarget Gene


$$$ΔΔCT = (CT(target,untreated) - CT(target,treated)) - (CT(ref,untreated) - CT(ref,treated))$$$
$$$ΔΔCT = (21.225 - 19.763) - (16.17 - 15.895)$$$
$$$ΔΔCT = 1.187$$$

$$$ΔΔCT = (CT(ref,treated) - CT(ref,untreated)) - (CT(target,treated) - CT(target,untreated))$$$
$$$ΔΔCT = (15.895 - 16.17) - (19.763 - 21.225)$$$
$$$ΔΔCT = 1.187$$$

$$$ΔΔCT = (CT(ref,untreated) - CT(ref,treated)) - (CT(target,untreated) - CT(target,treated))$$$
$$$ΔΔCT = (16.17 - 15.895) - (21.225 - 19.763)$$$
$$$ΔΔCT = -1.187$$$

$$$ΔΔCT = (CT(target,treated) - CT(target,untreated)) - (CT(ref,treated) - CT(ref,untreated))$$$
$$$ΔΔCT = (19.763 - 21.225) - (15.895 - 16.17)$$$
$$$ΔΔCT = -1.187$$$

As long as you stay consistent with the two ΔCT subtractions (i.e. always subtract treated from untreated, or reference gene from control gene, or vice versa) then the magnitude of the ΔΔCT part will always be the same.  The only thing that will change is the sign.

This is why you'll see some people will calculate the expression ratio as $$$2^{ΔΔCT}$$$ and others will do it as $$$2^{-ΔΔCT}$$$.  The difference is basically just how they set up their initial equation - effectively which direction they are comparing the samples.

To see why Delta Delta CT actually works we have to consider what's actually going on under the hood.

The absolute amount of material that we obtain through PCR for each sample for each primer pair is inversely proportional to $$$2^{CT}$$$.  We normalise our genes to a reference gene within each sample to ensure that we don't have any systematic errors due to differences between each sample (an internal control).

So the ratio of target gene to reference gene in each sample is therefore $$$2^{CT(target)} / 2^{CT(ref)}$$$.

However, because $$$\frac{b^c}{b^d} = b^{c-d}$$$ we can rewrite this as $$$2^{CT(target) - CT(ref)}$$$.

We then calculate the ratio between our two sample by calculating the quotient between the ratio of target gene to reference gene between the two samples as such:

$$$$Ratio = {2^{CT(target,untreated) - CT(ref,untreated)} \over 2^{CT(target,treated) - CT(ref,treated)}}$$$$

which by the same identity rule we applied before equals our ΔΔCT equation:

$$$(CT(target,untreated) - CT(ref,untreated)) - (CT(target,treated) - CT(ref,treated))$$$.


Standard Curve

An improvement to the Delta Delta CT method was introduced by Pfaffl to account for PCR efficiency curves deviating from the theoretical 100% efficient reaction.

To measure how efficient our PCR is for a given amplicon we run a template dilution series and see how closely our idealised PCR compares to real life.

If we run a dilution series of (0.25, 0.5, 1, 2) we would expect that there would be a one CT difference between each sample in the ideal 100% efficient reaction as shown:

CT ValueConcentration
360.25
350.5
341
332

However, if our actual measured CT values indicate a larger difference then our PCR reaction has been less efficient than we hoped.

CT ValueConcentration
360.25
34.90.5
33.81
32.72

We can work out exactly how much less efficient by comparing the CT values and the log of the Concentration.

We can do this on any log scale you like although commonly it'll be done on either $$$log_2$$$ or $$$log_{10}$$$ scales.  The result will come out the same either way.  We will use $$$log_2$$$ from now on since it fits well with the property of PCR doublings.



What we are interested in is the slope of the trend line.  In the case above the slope is -1.1.  We can then work out the efficiency of the reaction as

$$$Efficency = 2^{-1/slope} = 2^{-1/-1.1} = 1.878$$$

Therefore, for each PCR reaction we generate 1.878 copies of our template rather than, the theoretical, 2 copies in the ideal case.  The Efficiency is often then represented as a scale between 0-1 which is obtainable by subtracting 1 from the calculated efficiency above.

If we do a standard curve for each primer set we have (reference gene and target gene) then we can incorporate them into our Delta Delta CT equation to get:

$$$$Ratio = {Efficency(Target)^{CT(target,untreated) - CT(target,treated)} \over Efficency(Ref)^{CT(ref,untreated) - CT(ref,treated)}}$$$$

Note that the order of subtraction matters more here as we make sure that the the exponent of each calculated efficiency contains only the CT's which were produced by that primer pair.

ChIP-PCR

Unlike in RT-PCR for gene expression quantitation, ChIP-PCR will use a single primer pair per region of interest.  However, we will then usually also include an Input sample which the ChIP sample is compared with.

Input usually comprises a huge amount of DNA so it is usually necessary to take a subset of the Input as our actual sample to PCR.  The amount that is actually used may be between 1% and 10%.

To account for this we should apply an input adjustment.  This involves calculating a dilution factor which is equal to $$$1/ {Fraction Input}$$$.  For example if you have 1% input then your dilution factor (DF) is 1/0.01 = 100.  We then calculate our CT correction factor by calculating $$$log_{Efficiency}(Dilution Factor)$$$.

Worked example:  5% input is a DF of 1/0.05 = 20.  For the standard curve described above this is a CT correction of $$$log_{1.878}(20) = 4.75$$$.  This is then subtracted from each of the Input CTs before continuing as per the Standard Curve approach described above.

Because we only have a single primer pair we can use the same efficiency for the PCR throughout which allows us to simplify the Standard Curve approach to $$$Efficiency^{ΔΔCT}$$$.

Often you'll want to represent your result as a percent of input and this can be calculated for each condition as $$$Efficiency^{ΔCT}$$$.

Replicates & Error Propagation

One place that a number of mistakes seem to creep in is the treatment of replicates within PCR experiments.

The most frequent mistake I've seen is the use of the wrong type of mean when calculating the average ratio.  The key fact to remember is to always use the arithmetic mean on anything on a linear scale and always to use the geometric mean on anything on an exponential scale.

More concisely, if it's CT values (or differences thereof), then use an arithmetic mean.  If it's concentration values (anything that is $$$Efficiency^{CT}$$$) then use the geometric mean.

I also frequently encounter incorrect treatment of error bars.  Often people will take the standard deviation (or standard error) of concentration data, or ratios, and depict them directly in their bar charts.  However, these metrics assume normally distributed data which is only the case for the CT values themselves, not the ratios or concentrations.  Briefly, if your y axis on your chart is linear scale and your error bars are the same on both sides then this is wrong.

Finally we should deal with error propagation.  The error in your experiment is composed of the errors of $$$(Untreated,Ref) + (Untreated,Target) + (Treated,Ref) + (Treated,Target)$$$.  This error will usually be more than the error implied by taking the standard error of the ΔΔCT values for each replicate.  However, for uncorrelated variables we can propagate additive errors using the formula:

$$$Error(a+b) = \sqrt{ Error(a)^2 + Error(b)^2 }$$$

This works well for delta delta CT as the error is just the product of multiple additive errors.  However, for standard curve approaches we also have error from the Efficiency values we calculated for our curve and this can't be expressed as an additive factor.

Instead we need to use a Taylor series to propogate the error.  This allows for all six sources of variance (4x CTs + 2x Efficiencies) to be included in the final error calculation.

Control Gene Choice

One final factor is the choice of a suitable housekeeping gene as a control.  If you choose a reference housekeeping gene that changes wildly in between your untreated and treated samples then the results of your target gene compared to this gene will be wrong.

One approach to reduce the effect of this is to use the geometric mean of multiple housekeeping genes as your reference instead of choosing one.  Even still, it's important to carefully choose your reference genes for stability in your particular experiment.

Software

Another approach is just to use some of the software that's out there already.  One that's well used is REST 2009 which supports virtually anything you'll want to do.  Similarly there are open source solutions such as pyQPCR.  If you are willing to pay money you can also use something like GenEx.

If you do decide to make your own spreadsheet you should, at the very least, confirm that the answers you get from there are the same as the answers that you get from these programs.

I've uploaded a spreadsheet which illustrates most of the things I've talked about in this post.  It has examples for calculating standard curves, ΔΔCT, Standard Curve and Input Correction for ChIP-PCR.  It also has worked examples for error propagation for both ΔΔCT (via simple additive error propagation) and for Standard Curves (via a taylor series).

Realistically though, for almost all cases, and particularly anything rtPCR related, I recommend using something off the shelf like REST 2009.  There aren't many reasons you need to make your own spreadsheets - unless of course you are trying to explain how PCR normalisation works in a blog post.

Note that I've previously had errors in this spreadsheet myself.  I take this as a sign of just how easy it is to make these mistakes and also a reminder that there could well be errors still lurking in this sheet (and in any other spreadsheet or software you've used). Thanks most recently to Duncan Brian (UCL) for highlighting that the SE from the Taylor Series I took from REST384 was inversely correlated to the SE of the individual components. I've updated the sheet to include both the taylor series which matches the output of REST 384 and also the Taylor series described in the Gene Quantification Platform.

References

Gene Quantification Platform, Real Time PCR amended, A useful new approach? Statistical problems? http://www.gene-quantification.com/avery-rel-pcr-errors.pdf, Last Accessed 2017.

Livak, Kenneth J., and Thomas D. Schmittgen. Analysis of Relative Gene Expression Data Using Real-Time Quantitative PCR and the $$$2^{-ΔΔCT}$$$ Method. Methods 25.4 (2001).

Pfaffl, Michael W. A new mathematical model for relative quantification in real-time RT–PCR. Nucleic Acids Research 29.9 (2001).

Vandesompele, Jo, et al. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Genome biology 3.7 (2002).

Wednesday 22 May 2013

Undefined symbol (Rf_PrintWarnings) when using rpy 2.3.1 and R 3.0.1

So it seems the version of rpy2 (2.3.1) that comes with Ubuntu Raring isn't currently compatible with the version of R (3.0.1) in CRAN.

If you try to import it you'll get something like this:

$ python
Python 2.7.4 (default, Apr 19 2013, 18:28:01) 
[GCC 4.7.3] on linux2
Type "help", "copyright", "credits" or "license" for more information.
>>> import rpy2
>>> import rpy2.rinterface
Traceback (most recent call last):
  File "", line 1, in 
  File "/usr/lib/python2.7/dist-packages/rpy2/rinterface/__init__.py", line 93, in 
    from rpy2.rinterface._rinterface import *
ImportError: /usr/lib/python2.7/dist-packages/rpy2/rinterface/_rinterface.so: undefined symbol: Rf_PrintWarnings

However, the solution is pretty straightforward.

First download the source for rpy2 2.3.1. Then extract it and edit the file rpy2-2.3.1/rpy/rinterface/_rinterface.c. Then comment out the following on lines 1764-1765:
  extern void Rf_PrintWarnings(void);
  Rf_PrintWarnings(); /* show any warning messages */


Such that it appears as:
  //extern void Rf_PrintWarnings(void);
  //Rf_PrintWarnings(); /* show any warning messages */


Change to the rpy2-2.3.1 directory where you extracted the source to and reinstall rpy2 using:

sudo python setup.py install

Which should allow it to work again.
$ python
Python 2.7.4 (default, Apr 19 2013, 18:28:01) 
[GCC 4.7.3] on linux2
Type "help", "copyright", "credits" or "license" for more information.
>>> import rpy2
>>> import rpy2.rinterface
>>> 

You won't get any of the warning messages that Rf_PrintWarnings() would have output when it calls an R function.  This has already been fixed in the development branch but this quick fix means you don't have to wait for an official update to get rpy2 running again.

Thursday 18 April 2013

Placing the HIRA Histone Chaperone Complex in the Chromatin Landscape

Our lab is particularly interested in the HIRA chaperone complex made up of the HIRA, ASF1a, UBN1 and CABIN1 proteins. This post will summarise some of our latest work on the complex and some of the results we've found.

Background

DNA in cells is wrapped around a family proteins known as histones - this helps to compact the DNA to make it use up less space. There are four core histones (H2A,H2B,H3 and H4) as well as a linker histone (H1). Within those histones there are also a number variants. One we are interested in is the variant of histone H3 known as H3.3. This variant is different from other members of the H3 family (H3.1 and H3.2) in that it is produced all the time, not only in S-phase when the cells replicate. Our lab has an interest in cellular senescence which has strong links to both ageing and cancer sciences and the mechanisms that regulate histone integration in senescent (non-replicating) cells is crucial to understanding how they behave. Histone H3.3 has been implicated in the dynamics of nucleosome rearrangement and subsequently gene activation and ongoing transcription.

HIRA has previously been shown to be necessary for incorporating H3.3 into chromatin [1,2,3] outside of the replication cycle. A chaperone complex; consisting of HIRA, UBN1 and, as shown by our lab[4], CABIN1; cooperates with ASF1a (a histone-binding protein) to deposit H3.3 into chromatin. HIRA is required for embryo development as well as gene activation.

However, the distribution of the HIRA complex is unknown.

Genome Wide ChIP-Seq

To try to address this we did whole-genome ChIP-sequencing. Briefly, cells are crosslinked with a chemical which basically "sticks" proteins and DNA to each other. The resulting material is then sheared by sonication and incubated with antibodies that can specifically purify the protein of interest. Pieces of DNA that have been in contact with this protein at the time of crosslinking are co-purified in this process. We then sequence the resulting DNA fragments and align them back to a reference genome. We can then look at the genome and determine the frequency of that protein being located at that specific location in the genome.



We sequenced HIRA, UBN1, ASF1a. Comparing with a control input lane we found 8,296 regions with strong HIRA signals, 62,712 UBN1 regions and 64,550 ASF1a regions.



As you can see there's many more ASF1a and UBN1 regions than there are HIRA regions. This is not unexpected for two reasons: (1) HIRA is generally a harder protein to ChIP than ASF1a or UBN1 and therefore the complexity of the HIRA library was reduced compared to UBN1 and ASF1a libraries (2) both ASF1a and UBN1 are likely to be involved in other protein complexes that lack HIRA. At a minimum we've found the strongest HIRA peaks. We were interested in looking at the 1008 HIRA regions which have neither ASF1a or UBN1. Are they entirely because our criteria for our ASF1a and UBN1 regions are too strict?



We took our ASF1a dataset and steadily relaxed the criteria. Even relaxing the criteria we were unable to categorise much more than 93% of the HIRA as containing ASF1a (even when the FDR was essentially 1.0). Inspection reveals these regions have few or no reads for ASF1a or UBN1. This does at least rule out too stringent criteria as the reason for the existence of the HIRA-only regions.

Since we are mostly interested in the HIRA complex (rather than the other functions of ASF1a and UBN1) we concentrated on the HIRA regions. We compared our ChIP-seq results against publicly available ChIP-seq results in the ENCODE database and from Robertson et al.[5] and Vermeulen et al.[6] in the same cell line (HeLa) and hierarchically clustered the result (red = overlap, white = no overlap).


(Click for bigger).

From this clustering we were able to deduce the existence of four clusters (annotated above). The clustering is robust to different clustering linkages (ward, simple, complete etc) as well as addition or removal of factors.

Cluster 1 is enriched in H3K4me1, H3K4me3, H3K27ac, p300 and c-MYC, are FAIRE positive and DNase hypersenitive but are not within CpG islands, gene promotors and show little RNA polII. This suggests these are active enhancers for genes.

Cluster 2 is enriched in RNA polII, CpG islands, gene promotors, H3K4me3 and H3K27ac. This is consistent with these regions being promotors of actively transcribed genes.

Cluster 3 is enriched in H3K4me1 but does not have the same overlap with H3K27ac, p300 or c-MYC. They also lack overlap with promotors, CpG islands or RNA polII. This would likely classify them as inactive, weak or poised enhancers.

Cluster 4 is mostly HIRA-only regions (as compared to clusters 1-3 which have high overlap with both ASF1a and UBN1). It is also FAIRE and DNaseHS negative, not at gene promotors or CpG islands. Another important distinction of cluster 4 (HIRA-only) from clusters 1,2 and 3 (HIRA-complex) is the almost complete absence of H3.3.

As part of the same project we had also sequenced HA-H3.3 (measuring newly incorporated H3.3) and performed FAIRE-sequencing which are both shown in the heatmap above. If we look at these in more detail we can show the same clusters as before but plot the normalised read density around the HIRA peak. Also shown is ENCODEs H2Az dataset.


(Click for bigger).

Here we see results echoing the heatmap above but also showing the relative lack of reads in the 4th cluster (HIRA-only). You can also clearly see the nucleosome-free region at the transcription start site of genes in cluster 2 (gene promotors).

We looked at the gene promotors in more detail and showed that the HIRA complex was present at these nucleosome-free regions just upstream of the transcription start site.


(Click for bigger).

We also showed that the HIRA complex as well as H3.3 and H2Az are correlated with gene expression.


(Click for bigger).

Interactions

We then started looking at the proteins that were overlapping with HIRA. 76% of HIRA complex peaks (all three proteins) overlapped with at least one protein from four families:

* human SWI/SNF (BRG1, INI1, BAF155 and BAF170)
* AP-1 (c-FOS, c-JUN and JUND)
* c-MYC/MAX
* TFAP2 (TFAP2a and TFAP2C)

The strongest overlap (compared to chance) was with the SWI/SNF members which was particularly marked at active enhancers and promotors. As such this made them candidate HIRA complex binding partners.

To investigate this we tested interactions by immunoprecipitation-western blotting. We were able to show IP of endogenous HIRA also coprecipitates other members of the complex (UBN1, ASF1a) as well as transcription factors (c-JUN, c-MYC, GTF2i), SWI/SNF members (BRG1, BRM, INI1) and CTCF but not the negative controls of an abundant chromatin binding protein MCM2 and transcription factor TCF4. We also showed that the reverse was true of BRG1 and INI1 - that is; IP of BRG1 chromatin will also coprecipitate members of the HIRA-complex.


(Click for bigger).

We also performed a proximity ligation assay that scores close proximity of target proteins at the molecular level and showed that HIRA is located in close proximity to BRG1 and INI1 but not with proteins not known to interact with HIRA (DNMT1, MCM2, UACA, ATRX, XRN1, MBD2, LSH and EDC4).


(Click for bigger).

Discussion

We've done the first genome-wide study of human proteins HIRA, ASF1a and UBN1 in the chromatin and identified a number of likely binding partners. The SWI/SNF complex seems especially viable for further study in this respect. In fact just after this paper was accepted a former post-doc in the lab published that BRG1 was required for formation of Senescence-Associated Heterochromatin Foci (SAHF)[7] and he had also shown that HIRA was required SAHF[8] while in the lab. The cells we were working in were HeLa cells so we were unable to observe this effect in our current paper (HeLa cells are an immortal cell line so don't undergo replicative senescence in the usual way) but it seems likely that they may both be involved in the same process regulating SAHF formation.

The full paper can be freely accessed from Cell Reports[9] and the data can be downloaded from GEO GSE45025.

References

[1] Ray-Gallet, D., Quivy, J.P., Scamps, C., Martini, E.M., Lipinski, M., and Almouzni, G. (2002). HIRA is critical for a nucleosome assembly pathway independent of DNA synthesis. Mol. Cell 9, 1091–1100.

[2] Loppin, B., Bonnefoy, E., Anselme, C., Laurenc¸ on, A., Karr, T.L., and Couble, P. (2005). The histone H3.3 chaperone HIRA is essential for chromatin assembly in the male pronucleus. Nature 437, 1386–1390.

[3] Tagami, H., Ray-Gallet, D., Almouzni, G., and Nakatani, Y. (2004). Histone H3.1 and H3.3 complexes mediate nucleosome assembly pathways dependent or independent of DNA synthesis. Cell 116, 51–61.

[4] Rai, T.S., Puri, A., McBryan, T., Hoffman, J., Tang, Y., Pchelintsev, N.A., van Tuyn, J., Marmorstein, R., Schultz, D.C. and Adams, P.D. (2011) Human CABIN1 is a functional member of the human HIRA/UBN1/ASF1a histone H3.3 chaperone complex. Molecular and Cellular Biology.

[5] Robertson, A.G., Bilenky, M., Tam, A., Zhao, Y., Zeng, T., Thiessen, N., Cezard, T., Fejes, A.P., Wederell, E.D., Cullum, R., et al. (2008). Genome-wide relationship between histone H3 lysine 4 mono- and tri-methylation and transcription factor binding. Genome Res. 18, 1906–1917.

[6] Vermeulen, M., Eberl, H.C., Matarese, F., Marks, H., Denissov, S., Butter, F., Lee, K.K., Olsen, J.V., Hyman, A.A., Stunnenberg, H.G., et al. (2010). Quantitative interaction proteomics and genome-wide profiling of epigenetic histone marks and their readers. Cell 142, 967–980.

[7] Tu, Z., Zhuang, X., Yao, Y.G. and Zhang, R. (2013) BRG1 Is Required for Formation of Senescence-Associated Heterochromatin Foci Induced by Oncogenic RAS or BRCA1 Loss. Mol Cell Biol.May;33(9):1819-29.

[8] Zhang, R., Poustovoitov, M.V., Ye, X., Santos, H.A., Chen, W., Daganzo, S.M., Erzberger, J.P., Serebriiskii, I.G., Canutescu, A.A., Dunbrack, R.L., Pehrson, J.R., Berger, J.M., Kaufman, P.D., Adams, P.D. (2005) Formation of MacroH2A-containing senescence-associated heterochromatin foci and senescence driven by ASF1a and HIRA. Dev Cell. 2005 Jan;8(1):19-30.

[9] Pchelintsev, N.A., McBryan, T., Rai, T.S., van Tuyn, J., Ray-Gallet, D., Almouzni, G. and Adams PD. (2013) Placing the HIRA Histone Chaperone Complex in the Chromatin Landscape. Cell Reports.

Friday 5 April 2013

Minimal UCSC Track Hubs

I frequently display results within the Lab using the UCSC genome browser and I'm often asked how I do the "overlapping tracks" feature - specifically I'm usually asked for a minimal configuration that they can then extend.



It's something UCSC has been doing on their own tracks for some time (e.g. the Promotor/Enhancer Histones mark track) but was made available to 3rd parties reasonably recently through their track hub feature.

The first requirement is that you will have to have somewhere you can upload bigWig files to that UCSC can access via HTTP. You can generate bigWig files using UCSC apps bedGraphToBigWig or wigToBigWig.

Note that if you intend on displaying tracks related together they should be appropriately normalised to make comparisons make sense. If your data is already normalised (for example % methylation from Bisulfite sequencing) then this is fine. Otherwise, if you have a bedGraph file representing a pileup (number of reads covering each basepair) then something like the following should work to normalise by library size:

awk -v NUM_FRAGS=x 'BEGIN{OFS="\t"}{print $1,$2,$3,$4/(NUM_FRAGS/1000000)}' file.bedGraph > file.normalised.bedGraph
Where NUM_FRAGS=x is changed to the correct number of fragments in your dataset. At this point you can follow the instructions from UCSC to setup your first track hub. I will briefly summarise them here though.

There's a number of little files required to make it work right.. The file structure I use is something like this:
/hub.txt
/genomes.txt
/hg18/trackDb.txt
/hg18/project1.txt
Here hub.txt is the "root" of the hub and is the URL that you give to UCSC to initialise the hub and load it into the browser. The format is minimal and just describes the human readable description of your hub and the location of a "genomes" file.

/hub.txt
hub HubNameHere
shortLabel Label
longLabel Longer Label Here
genomesFile genomes.txt
email tony@mcbryan.co.uk
The genomes file is simply a list of each genome you want to be able to display data against. Valid options are any of the genomes that UCSC supports on the browser (or your mirror).

Minimally this must include one genome and link to the associated trackDB file which describes the tracks to be loaded.

/genomes.txt
genome hg18
trackDb hg18/trackDb.txt
UCSC uses trackDb files internally to format the display of tracks onto the genome. Track Hubs allow us to use a subset of the trackDb functionality to load our own tracks. UCSC has long had a Custom Tracks feature but this only ever allowed a very small fraction of the features of the trackDb while the Track Hubs allow us much more.
You can, if desired, simply keep all your track definitions in the single trackDb file but this will quickly get messy. Luckily we can import additional files.

/hg18/trackDb.txt
include project1.txt
While in project1.txt we can include only the track definitions that are relevant to that project.

/hg18/project1.txt
track SuperTrack
shortLabel Label
longLabel Long Label
superTrack on none
priority 1

track CompositeTrack
container multiWig
configurable on
shortLabel Label
longLabel Long Label
visibility hide
type bigWig 0 100
autoScale off
aggregate overlay
viewLimits 0:100
windowingFunction mean
superTrack SuperTrack full
showSubtrackColorOnUi on

track SubTrack1
type bigWig
shortLabel Label
longLabel Long Label
parent CompositeTrack
visibility full
bigDataUrl http://a.b.com/path/to/wigs/SubTrack1.bigWig
color 51,102,153

track SubTrack2
type bigWig
shortLabel Label
longLabel Long Label
parent CompositeTrack
visibility full
bigDataUrl http://user:pass@a.b.com/path/to/wigs/SubTrack2.bigWig
color 255,102,0
Therefore project1.txt is where the magic happens. We have a super track (named SuperTrack) which contains all other tracks. This gives us a single drop-down box on the UCSC browser which we can use to turn a whole projects tracks on or off at once.

This super track contains a single composite track which is of type multiWig. The multiWig is capable of overlaying multiple wig tracks on top of each other (using the aggregate option). To make this work we create regular bigWig tracks with their parent set to the composite track.

The bigDataUrl is the important one as that tells UCSC how to actually get to your data (hosted on a web server somewhere). Note that UCSC supports providing the URL along with username and password (as shown above in SubTrack2). This is compatible with usernames and passwords enforced by htaccess on Apache server.

You can fix the viewlimits on wig files (here fixed at 0 to 100 within the composite track. Note that settings on a parent track are inherited by their child tracks unless the child has a conflicting setting.

If you prefer you can have a dynamic range (may be appropriate for ChIP-Seq data etc) which will resize the view to the bounds of your data. If you use dynamic scaling you probably also want to include the alwaysZero attribute on the composite track and set it to "on" otherwise UCSC will scale from your lowest value to your highest value within the window.

You can then view the track hub by entering it on the UCSC Hub Connect page or by creating a link such as:

http://genome.ucsc.edu/cgi-bin/hgTracks?db=hg18&hubUrl=http://user:pass@a.b.com/hub/hub.txt

You should be able to use the above as a reference point to building your own Track Hubs. For more information, and details on the other features you can use in Track Hubs, see UCSC's help pages.

Wednesday 27 February 2013

The Binomial Distribution, Python and Bisulphite Sequencing

This is Part 2 in a series on Bisulphite Sequencing.
You can return to Part 1 (Post Processing Bismark Bisulphite Sequencing Data) or skip to Part 3 (Simple Visualisation of Bisulphite Sequencing Data).


Today I'm going to describe how the binomial distribution can be applied to bisulphite sequencing data in order to accurately determine the number of CpG sites which exhibit methylation in a sample.

The binomial distribution is actually pretty straightforward.  If you have a coin that has a probability $$$p$$$ to come up heads then the binomial distribution can tell us the probability that if you throw $$$n$$$ coins then what is the probability that you will get $$$k$$$ heads.  We call this the probability mass function.

Let's do the simplest example first.

If I throw one coin and it has a probability of 0.5 (an unbiased coin) then we have two outcomes; we throw a heads or we throw a tails.


Result of 1st Coin Toss Number of Heads Overall Probability
Tails (p=0.5) 0 0.5
Heads (p=0.5) 1 0.5


Ok, that's fairly trivial. We have two possible outcomes each with a possibility of 0.5. So our probability of getting exactly one heads is simply 0.5.

Let's add a second coin toss. Again we keep our probability as 0.5 chance for a heads.


Result of 1st Coin Toss Result of 2nd Coin Toss Number of Heads Overall Probability
Tails (p=0.5) Tails (p=0.5) 0 $$$0.5 * 0.5 = 0.25$$$
Tails (p=0.5) Heads (p=0.5) 1 $$$0.5 * 0.5 = 0.25$$$
Heads (p=0.5) Tails (p=0.5) 1 $$$0.5 * 0.5 = 0.25$$$
Heads (p=0.5) Heads (p=0.5) 2 $$$0.5 * 0.5 = 0.25$$$

Each coin toss still has a 0.5 probability of coming up heads but we now have four possible results.  We can work out the probability of each result by multiplying the probability of the first coin toss by the probability of the second coin toss ($$$0.5 * 0.5 = 0.25$$$ for each row above).

Note that in both examples above the probabilities for the table always add up to 1.  In the second table we have one row where we get 2 heads, two rows where we get 1 heads and one row where we get 0 heads.  So our cumulative probability for getting 2, 1 or 0 heads is 0.25, 0.5 and 0.25 respectively.

Let's change the odds.  We now have a coin that we've fixed to be strongly biased towards tails.  Our probability is now only 0.05 that we will get a heads and 0.95 that we will get a tails.

Result of 1st Coin Toss Result of 2nd Coin Toss Number of Heads Overall Probability
Tails (p=0.95) Tails (p=0.95) 0 $$$0.95 * 0.95 = 0.9025$$$
Tails (p=0.95) Heads (p=0.05) 1 $$$0.95 * 0.05 = 0.0475$$$
Heads (p=0.05) Tails (p=0.95) 1 $$$0.05 * 0.95 = 0.0475$$$
Heads (p=0.05) Heads (p=0.05) 2 $$$0.05 * 0.05 = 0.0025$$$

We expect in most cases that we will get two tails and we see that that happens about 90% of the time from above.  We can also see that our probability of getting two heads is the sum of the middle two rows -- but that each has a low probability.  Nonetheless we have a 9.5% of getting at least one heads from either coin toss ($$$0.0475+0.0475=0.095$$$). Bringing up the rear is a very small, 1/4 of a percent, chance that we get two heads.

We don't want to work out a table like above every time we want to know the probability, so, we can define a function as follows:

$$$$binom.pmf(k,n,p) = {\frac{n!}{k!(n-k)!}}p^k(1-p)^{n-k}$$$$
Where $$$k$$$=number of heads, $$$n$$$=number of coin tosses and $$$p$$$=probability of heads.  The $$${\frac{n!}{k!(n-k)!}}$$$ portion of the function above is simply the binomial coefficient or the "n choose k" function (i.e. "how many different ways can this happen"). You'll often find this on a calculator as $$$_nC_k$$$. For more details on how this function works see the Wikipedia article on the Binomial Distribution.  Otherwise, observe that:

$$$$binom.pmf(0,2,0.05) = 0.9025$$$$
$$$$binom.pmf(1,2,0.05) = 0.095$$$$
$$$$binom.pmf(2,2,0.05) = 0.0025$$$$

Luckily, we don't need to implement the distribution ourselves as most programming languages will have a stats library with this function in it somewhere.  In Python this is:

>>> import scipy.stats
>>> scipy.stats.binom.pmf([0,1,2],2,0.05)
array([ 0.9025,  0.095 ,  0.0025])

Excellent, that matches exactly what we had above.  Now to a practical application.

In bisulphite sequencing we sequence lots of CpG's and the result is either a C or T (methylated or unmethylated respectively).  We essentially have a large matrix representing the number of methylated and unmethylated reads which will look like the following (entirely made up data).

chr1    1000    5    2
chr1    1025    4    0
chr1    1026    1    10
chr1    1055    1    0
chr1    1089    0    1
chr1    1090    1    1

There are a number of places errors can occur but in particular we are worried about seeing methylated signal where no such signal exists.  This can happen for a number of reasons; including an incorrectly sequenced base (sequencing error) or incomplete bisulphite conversion.

If we did our sequencing properly we also sequenced an unmethylated spike-in control (usually lambda).  As a result we can work out an overall error rate by seeing how many methylated bases we sequence from our unmethylated control.  Divide this by the total number of (relevant) bases sequenced from our control and we have our overall probability of reading a C (methylated) when it was actually a T (unmethylated).

For each CpG we sequenced we want to estimate the probability (p-value) that we would have sequenced that number (or more) methylated bases under the assumption that the CpG in question was actually entirely unmethylated (our null hypothesis) within the cell population.

If we sequence any given CpG and find that 4 out of 10 reads are methylated, and that our error rate from our lambda sequences was 0.005, then we can calculate the probability of this happening by summing the binom.pmf for each result.  i.e. What is the probability that we would get either 4,5,6,7,8,9 or 10 methylated bases out of 10 total reads based on errors alone. The number of methylated bases is used in place of the number of heads and the number of sequenced bases is used instead of the number of coin tosses.  The error rate can then equivalent to the probability of throwing a heads.

>>> import math
>>> math.fsum(scipy.stats.binom.pmf([4,5,6,7,8,9,10],10,0.005))
1.2813262561414168e-07

We can formalise this as the cumulative distribution function which calculates the sum of the probability mass function from $$$0...k$$$ (still parametrised by $$$n$$$ and $$$p$$$).  Since all the probabilities must add to 1 then the probability of 4 or more heads is equal to 1 minus the probability of 3 or less heads (0,1,2 or 3).

$$$$binom.cdf = \sum_{i=0}^{k} {\frac{n!}{i!(n-i)!}}p^i(1-p)^{n-i}$$$$

However, the way we are working this out isn't ideal.  We introduce float precision errors for every pmf calculated as well as another precision error when we actually sum them.  (Although math.fsum does reduce this final error over a naive sum by tracking the cumulative error from intermediate partial sums). Luckily we can avoid summing the terms directly and use an implementation based on the incomplete beta integral as available in the python scipy package which should be higher precision:

>>> 1-scipy.stats.binom.cdf(4-1,10,0.005)
1.2813262562083594e-07

Note above that we calculate the cumulative distribution function of $$$k=4-1$$$ (i.e. 3) so we get the cumulative probability for 0,1,2 or 3 heads. 

The survival function is the same as above but instead of operating over the range $$$0...k$$$ it works for the range $$$(k+1)...n$$$.

$$$$binom.sf = \sum_{i=k+1}^{n} {\frac{n!}{i!(n-i)!}}p^i(1-p)^{n-i}$$$$
Which in scipy is:
>>> scipy.stats.binom.sf(4-1,10,0.005)
1.2813262561414144e-07

Note again that we use $$$k-1$$$.  This time it is because the survival function works from $$$(k+1)...n$$$ so if we want the range $$$k..n$$$ then we need to subtract 1 from $$$k$$$.

All of the above tests are one-tailed (probability of equal to or greater number of heads).  If we want to test for divergence from the expected number of heads (either greater than or less than) we want to use a two-tailed test which splits the probability across the two tails.

To calculate a two-tailed p-value we use the method of small p-values which sums all the p-values which are smaller than the probability of getting exactly $$$k$$$ out of $$$n$$$ heads.

def binom_test(k,n,p):
    # probability of exactly this result
    exactp = scipy.stats.binom.pmf(k,n,p)
    
    # p-vals for every possibility
    allpvalues = scipy.stats.binom.pmf(range(0,n+1),n,p)
    
    # p-vals smaller than exactp
    smallpvalues = [x if x <= exactp else 0 for x in allpvalues]
    
    return math.fsum(smallpvalues)

However, again we have the issue that we are calculating lots of individual probability mass functions and adding the results which will introduce small precision errors.  Luckily, scipy has a more optimised version of this function which uses the cumulative distribution and survival functions to increase the precision.

>>> scipy.stats.binom_test(4,10,0.005)
1.2813262561414144e-07

In the example above the expected number of heads ($$$p*n$$$) is very low so using two-tailed p-values doesn't change the p-value.  This is because the p-values are essentially ordered from largest to smallest for the range $$$0...n$$$ as the most likely value for $$$k$$$ under the null hypothesis is 0 with increasingly smaller probabilities as $$$k$$$ is increased.  As such the only p-values that contribute are those which are in the "greater than" tail (the range $$$k...n$$$).

However, if the error rate is higher then the choice of one-tailed vs two-tailed will obviously make more of a difference.

# one tailed test
>>> scipy.stats.binom.sf(10-1,20,0.3)
0.0479618973313434

# two-tailed test
>>> scipy.stats.binom_test(10,20,0.3)
0.08344502962981204

Now we can use the binom_test function from scipy to implement a methylation caller in Python.  This will calculate the probability that the methylation we see at any given CpG is purely the result of errors.  We use the same file format as in the previous post such that the first two columns identify the CpG and the remaining columns are pairs of methylated and unmethylated counts for a number of samples.

import getopt
import sys
import csv
from scipy.stats import binom_test
from itertools import izip_longest
from multiprocessing import Pool

def binomRow(row):
    chrm, coord, data = row

    outputRow = [chrm, coord]
    for rowextension in map(binomMethCall,data):
        outputRow.extend(rowextension)
    return outputRow

def binomMethCall(group):
        meth,unmeth,errorrate = group
        return [meth,
                unmeth,
                binom_test(meth,meth+unmeth,errorrate)]

def grouper(n, iterable, fillvalue=None):
    "grouper(3, 'ABCDEFG', 'x') --> ABC DEF Gxx"
    args = [iter(iterable)] * n
    return izip_longest(fillvalue=fillvalue, *args)

def unpackLine(chrm,coord,*data):
    return chrm.strip(), int(coord), [int(d) for d in data]

def extractLine(line):
    chrm,coord,data = unpackLine(*line)
    if not chrm.startswith("chr"):
        chrm = "chr"+chrm
    return (chrm,coord,data)
    
if __name__ == '__main__':
    try:
        opts, args = getopt.getopt(sys.argv[1:], "", ["combined=","errorRates=","cores="])
    except getopt.GetoptError, err:
        # print help information and exit:
        print str(err) # will print something like "option -a not recognized"
        sys.exit(2)
    
    infile = None
    cores = 4
    
    for o, a in opts:
        if o=="--combined":
            infile = a
            print "combined", a
        elif o=="--errorRates":
            errorRates = [float(x) for x in a.split(",")]
            print "Error Rates", errorRates
        elif o=="--cores":
            cores = int(a)
    
    assert infile != None
    
    methfile = csv.reader(open(infile, "r"), delimiter="\t")
    
    binomOutput = csv.writer(open(infile+".binom","w"),delimiter='\t')

    p = Pool(cores)
    
    rows = []    
    for count,line in enumerate(methfile):
        
        if count % 100000 == 0:
            print count
            for row in p.map(binomRow,rows):
                binomOutput.writerow(row)
            rows = []
        
        chrm,coord,data = extractLine(line)
        
        row = [chrm,coord]
        
        # extract in pairs
        samples = [(meth,unmeth,errorRates[sampleno]) for sampleno,(meth,unmeth) in enumerate(grouper(2,data))]        
        row.append(samples)
        rows.append(row)
    
    print count
    for row in p.map(binomRow,rows):
        binomOutput.writerow(row)
        
        

Executed as:

python binomialMethCall.py \
--combined "Samples.1.and.2.CpG_context.aggregate" \
--errorRates 0.005,0.002 

Where the error rates are obtained from the lambda spike-in alignments for each sample.  This will create an extra column per sample.  You should correct the p-values using the False Discovery Rate method and threshold at at a suitable value (FDR<=0.01).

Update: (12 July 2013): I rewrote a chunk of the binomialMethCall python script to take better advantage of pythons multiprocessing libraries. The updated version should be better at maxing out the number of cores you requested it to use.