Thesis AI - Carlos Fenollosa Bielsa

Download as pdf or txt
Download as pdf or txt
You are on page 1of 89

Universitat Politècnica de Catalunya

Departament de Llenguatges i Sistemes Informàtics


Màster en Intel.ligència Artificial

Master Thesis

Using AI techniques to determine promoter


location based on DNA structure calculations

Carlos Fenollosa Bielsa

Advisors:
Ramon Goñi Macià, MMB, PCB
Javier Vázquez-Salceda, LSI, UPC

Barcelona, September 2008


“To build computers that are like humans, we first need to understand how
we were designed. Bad thing, as we were never designed, but evolved from the
chaos with no documentation left.”
Abstract

DNA sequencing projects have started the race to fully annotate complete
genomes, including the human one. Despite that, little is known about ge-
netic regulation, the mechanisms that control where and when the genes are
expressed, and promoters are maybe the most important of these mechanisms.
An increasing number of studies have been focused on the DNA molecule
and its structure. This has lead to a set of physical properties which can be com-
puted from mathematical models, and describe some aspects of this molecule.
Unfortunately, the existing tools are scattered through the different web sites of
many research groups, and extracting data with them is still very unpleasant.
The first part of this thesis presents DNAlive, a new platform to calculate DNA
physical properties, showing the results in a visual and useful way for genetic
researchers, cross-linking the data with external databases.
For the second part, a full study of DNA physical descriptors has been
performed, revealing significative similarities between them. Using that data,
a set of neural networks has been developed to detect promoters on a DNA
sequence. The resulting software is the second version of ProStar, the MMB
group’s1 latest promoter predictor.

1 The Molecular Modelling & Bioinformatics Group (http://mmb.pcb.ub.es) is a research

group hosted at the Parc Cientı́fic de Barcelona. I have been working there since 2007, in a
joint programme with the Barcelona Supercomputing Center, and this MSc Thesis has been
developed as one of the group’s projects.
Acknowledgements

This Thesis could be summed up as Structural biology and genetics for com-
puter scientists: From nothing to “a lot” in just a year. That is why these
acknowledgements are the most important part of this thesis.
First of all, I need to thank Ramon Goñi for offering me a career in the
Life Sciences, for his invaluable help and guidance since day one, for his interest
and dedication, for calling me from Madrid whenever it was needed and sending
me e-mails at late night hours. You reviewed my work as if it were your own
responsibility. Thanks for that.
On the other side, I’m sure that Javier Vázquez has also learned a lot of
biology by brute force. It’s not easy to understand genes and promoters, let
alone if the teacher is me. Everything would have been easier if I had chosen
the blue pill named Contract, but I love challenges. Thanks, and sorry!
I’d also like to thank Modesto Orozco for getting time from nowhere to meet
with me and teach me concepts from Biology 101. Maybe he would get the
Nobel Prize if he finally admitted that he has discovered 30-hour-long days.
At the LSI, Karina Gibert and Lluı́s Belanche offered me great help on the
PCA and the neural network. I still keep the napkin from the cafeteria where
Karina drafted the matrix I needed to build to run the PCA.
Thanks to José Antonio Alcántara for his fast help with all my server issues.
A special acknowledgement also for the people at the INB, and I mean all of
them, for explaining me the internals of BioMoby, which are not always nice.
Many eyes have been upon DNAlive and have helped me to debug the whole
platform, so thanks to the people in the MMB group in general and the web
testers in particular, especially David Torrents at the BSC.
To all my workmates, I’m glad I’ve met you! It is a pleasure to work with
you and a privilege to have those profound conversations from time to time.
In no particular order, but ladies first: Montse, Laura, Jordi and Felix, with
occasional outsiders.
Finally, as I don’t want to leave anybody out of these acknowledgements,
here’s a placeholder: thanks to $you2 too!

But wait, there’s more!

To my mom: it’s okay to be stressed when there’s only a week left to hand
the manuscript in. Don’t try to calm me down, I need the stress to stay awake.
You are guilty for teaching me curiosity and ambition, paying my University so
that I could study on weekends, I don’t want to screw it up in the last minute.
2 PHP joke. Not Perl. PHP.

i
I’m fine now. I’m sure that grandpas will be very happy because I have finished,
at last.
To my friends: It’s okay not to go out saturday nights if you need to stay
until five o’clock AM running batches in the computer nodes. I don’t want to
disconnect and relax for a couple of hours because I’ll feel guilty afterwards, but
I want to tell you that I appreciate the effort. Really. Remember how you were
so worried that you didn’t allow me to attend your PFC defense? We can party
afterwards (and we will).
This Thesis also accounts as my PFC for Computer Engineering. Seven long
years, the latter two attending the MSc, have allowed me to meet a lot of cool
people at the FIB. Some carried on, some dropped out, some switched to the
technical engineering and finished in only three years. As I love writing —I
think you have already noticed it by the length of this section— on day zero
I joined l’Oasi, a student group at the FIB that publishes a magazine every
semester. This was the best decision I made during my studies, because I found
the boldest people in college there. l’Oasi has been my second home for five
years.

But wait, there’s even more!

The acknowledgements finish here, however. If you want to keep reading,


you will find a personal analysis on everything regarding my experience on this
thesis. Not the scientific conclusions, not textbook-like information, but my
own experience. I suppose that I could have posted it to my blog instead, but
I want it attached to this document. You have been warned.
When you do blind data mining, you learn something: data is just data.
The trick is how to interpret it. But how is a computer engineer supposed to
discover whether the correlation between something named “a-philicity” and
“propeller twist” looks good or not?
Only after a year of working in bioinformatics I am able to slightly under-
stand the data I’m working with. For so many time it was just numbers that
only took shape when shown to my advisors. I was working blindly. I could only
discover that the algorithm for calculating correlations had a bug when Ramón
told me that the results “didn’t look good”.
Working in science has something special. You could get disastrous results
only to discover later that the methods are OK but there is a bug in a small
module of your software. And sometimes, even published papers contain erro-
neous data that will be refuted years later. Science is about discovering things,
things that nobody else but you is working on. Results that nobody can tell
whether they are correct or not, because everything is so new. Science is not
math, biology is not a computer algorithm which will output a one or a zero.
I thought that neural networks are nice because they almost do magic. They
can classify anything, and sometimes pretty good. Unfortunately, that turns out
to be true only for the Iris.arff3 dataset and a couple more which a toddler could
linearly separate with a crayon and a ruler. In biology, data has a lot of noise,
the datasets might be not 100% trustworthy and we don’t even know how to
start modeling what happens in our body. Did I mention that we can extract
huge amounts of information that can’t be treated computationally?
3 Machine-learning joke

ii
As an example, the gene expression is so tricky that there could be a gene
hidden in some obscure region of the DNA and we don’t know of its existence
because it expresses only in the liver, and only in the first month of pregnancy.
Then, it becomes silent forever. How are we supposed to discover it? Do we
need to make a map of all the expressed DNA in all organs for every 10 minutes
since the first second of gestation? The answer is even worse than the question:
yes, we need; and no, we don’t know how to generate this data. Or where to
store and process it.
But it turns out that a computer guy can learn biology. Fortunately, all the
institutions that form the MMB group have weekly seminars —this sometimes
means more than five seminars a week— where everybody can learn from every
field. Some are boring, some are interesting, some are so specific that you
only understand them some months later, after working a bit on that field and
reading a lot of Wikipedia articles4 . But, in the end, you learn something.
It’s funny because biologists and chemists also tell you their own challenge:
learn to use a computer and program software. They might have a PhD, but
some only know enough computer science to run MS Word in Windows, because
they’ve always worked in a lab. The solution? Collaborate. They explain me
what a propeller is, and I explain them how to Perl5 . Sadly, from time to time I
am asked a question on FORTRAN, and I can’t help. In the end, it is all about
mixing experiences, and it is very rewarding.

I studied computer engineering because I wanted to know more about com-


puters. And, thankfully, when the major is finished, computers have no secrets.
One can explain the whole process between pressing a key on a laptop and get-
ting a map from Google, step by step. Unfortunately, we can explain less things
about men, society and the universe. We need knowledge, collaboration, com-
puters, and further developments in Artificial Intelligence. Nowadays, a giant
leap in computer architectures is required, because even with huge artifacts like
the Mare Nostrum, Turing Machines are basic sequential machines which are
not that good for processing real world data.
That being said, if you have reached this end line, the final acknowledgement
goes for you. Thanks for reading.

4 How could we live without Wikipedia?


5 Itseems that a lot of bioinformatic libraries are written in Perl. I also invest some time
on trying to convince them to switch to PHP for their own scripts, with mixed results.

iii
Contents

1 Introduction 2
1.1 Definition of the problem and objectives . . . . . . . . . . . . . . 3
1.2 Application of the AI . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Structure of this thesis . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Biology 6
2.1 Central dogma of the molecular biology . . . . . . . . . . . . . . 6
2.2 Genes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Promoters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4 Physical properties of the DNA . . . . . . . . . . . . . . . . . . . 12
2.5 Nucleosomes and Chromosomes . . . . . . . . . . . . . . . . . . . 15
2.6 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . 15

3 State of the art 17


3.1 Genome browsers . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Sequence search . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3 DNA dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.4 Gene predictors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5 ProStar 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.6 EP3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.7 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . 25

4 Creating a platform for the analysis of DNA: DNAlive 26


4.1 Objective and requirements . . . . . . . . . . . . . . . . . . . . . 26
4.2 System architecture and design . . . . . . . . . . . . . . . . . . . 27
4.3 The physical properties scripts . . . . . . . . . . . . . . . . . . . 27
4.4 Web page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.5 Web services . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.6 Deployment and testing . . . . . . . . . . . . . . . . . . . . . . . 34
4.7 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . 34

5 Applying AI techniques to ProStar 36


5.1 Proposed solution . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.2 Promoter division in four groups . . . . . . . . . . . . . . . . . . 37
5.3 DNA descriptor analysis . . . . . . . . . . . . . . . . . . . . . . . 38
5.4 Predictor training . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.5 ProStar 2.0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.6 Evaluation of the results . . . . . . . . . . . . . . . . . . . . . . . 44

iv
5.7 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . 44

6 Results 47
6.1 DNAlive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.2 ProStar 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

7 Conclusions 56
7.1 DNAlive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.2 ProStar 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.3 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

A DNAlive publication 67

B Description of the physical properties 73


B.1 Unusual DNA conformation . . . . . . . . . . . . . . . . . . . . . 74
B.2 DNA disruption energy . . . . . . . . . . . . . . . . . . . . . . . 75
B.3 DNA 3DNA structure . . . . . . . . . . . . . . . . . . . . . . . . 76
B.4 DNA flexibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
B.5 DNA stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
B.6 DNA non-linear dynamics . . . . . . . . . . . . . . . . . . . . . . 78
B.7 PARMBSC0 Helical force constants . . . . . . . . . . . . . . . . . 78
B.8 Regulation parameters . . . . . . . . . . . . . . . . . . . . . . . . 79

v
List of Figures

2.1 Cell types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6


2.2 A chunk of DNA in its stable 3D form . . . . . . . . . . . . . . . 7
2.3 DNA chemical structure . . . . . . . . . . . . . . . . . . . . . . . 8
2.4 Central dogma of the molecular biology . . . . . . . . . . . . . . 9
2.5 The process of gene expression . . . . . . . . . . . . . . . . . . . 11
2.6 The structures of A-DNA, B-DNA and Z-DNA . . . . . . . . . . 13
2.7 A triplex structure. . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.8 A quadruplex structure. . . . . . . . . . . . . . . . . . . . . . . . 14
2.9 A nucleosome bound to the DNA. . . . . . . . . . . . . . . . . . 15
2.10 From the DNA string to the chromosome. . . . . . . . . . . . . . 16

3.1 The UCSC Genome Browser. . . . . . . . . . . . . . . . . . . . . 18


3.2 Managing tracks in the UCSC Genome Browser. . . . . . . . . . 18
3.3 ProStar1: Calculus of the force constants . . . . . . . . . . . . . 22
3.4 ProStar1: Generation of the promoter database . . . . . . . . . . 22
3.5 ProStar1: System architecture . . . . . . . . . . . . . . . . . . . . 23
3.6 Physical properties study for EP3 . . . . . . . . . . . . . . . . . . 24

4.1 Architecture of DNAlive . . . . . . . . . . . . . . . . . . . . . . . 28


4.2 Plot of the “A-Philicity” property for a given sequence . . . . . . 29
4.3 Physical properties web services . . . . . . . . . . . . . . . . . . . 32
4.4 Structural web services . . . . . . . . . . . . . . . . . . . . . . . . 32

5.1 Descriptor analysis methodology. . . . . . . . . . . . . . . . . . . 40


5.2 Data models for the predictor training. . . . . . . . . . . . . . . . 43
5.3 ProStar2 algorithm. . . . . . . . . . . . . . . . . . . . . . . . . . 45

6.1 Screen shot of a DNAlive session . . . . . . . . . . . . . . . . . . 48


6.2 Plot of the signal for the CpG property . . . . . . . . . . . . . . 49
6.3 Plot of the signal for the Curvature property . . . . . . . . . . . 50
6.4 Pearson correlation matrix for the TATA+ CpG- group. . . . . . 52
6.5 Weka plotting the correlation between Stability and Duplex Sta-
bility Free Energy. . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.6 Screen shot of a ProStar 2 session . . . . . . . . . . . . . . . . . . 55

vi
List of Tables

4.1 Sample values for different physical properties . . . . . . . . . . . 29

5.1 Weight matrix for the TATA box . . . . . . . . . . . . . . . . . . 38


5.2 Elements in the dataset . . . . . . . . . . . . . . . . . . . . . . . 38
5.3 Sample descriptors matrix I before applying the eigenvectors
function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.4 Elements per group for the predictor training . . . . . . . . . . . 42

6.1 Resulting eigenvectors after the PCA . . . . . . . . . . . . . . . . 51


6.2 The first three eigenvectors extracted from the tata-cpg+’s PCA 53
6.3 Neural network models for each promoter group . . . . . . . . . . 54
6.4 Accuracy results for ProStar 2 versus ProStar 1 . . . . . . . . . . 54

B.1 Unusual DNA conformation . . . . . . . . . . . . . . . . . . . . . 74


B.2 DNA disruption energy . . . . . . . . . . . . . . . . . . . . . . . 75
B.3 DNA 3DNA structure . . . . . . . . . . . . . . . . . . . . . . . . 76
B.4 DNA Flexibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
B.5 DNA stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
B.6 DNA non-linear dynamics . . . . . . . . . . . . . . . . . . . . . . 78
B.7 PARMBSC0 Helical force constants. . . . . . . . . . . . . . . . . 78
B.8 Regulation parameters . . . . . . . . . . . . . . . . . . . . . . . . 79

1
Chapter 1

Introduction

Sequencing projects have provided the genome sequence of many evolved organ-
isms, including mammals. Unfortunately, less information exists on the detailed
mechanisms controlling gene expression.
Promoter1 and gene detection is one of today’s biggest bioinformatic prob-
lems because of its complexity. Genes control all the functions of the human
body, but the models found to date do not work properly for some kind of genes.
Furthermore, it is a recent field of study, and very important research groups are
trying to tackle this problem with bioinformatic tools. The use of AI techniques
over statistical models can help researchers to find genetic models, even if they
cannot be fully understood.
Despite the impressive success of high-throughput experimental techniques,
the determination of promoters is still a big challenge in the post-genomic era.
Promoter prediction would be straightforward if all genes were perfectly anno-
tated and promoters were always placed at the same relative location respective
to the gene.
Unfortunately, annotation is often subject of large uncertainties and recent
genomic analysis has shown that the naive concept of a gene having a single start
point located near the gene is not correct. Furthermore, one gene region might
have several Transcription Start Sites, one promoter might induce transcription
start at different sites and promoter regions often overlap [Cea05]. On the other
hand, sequence signals associated to promoters are rather unspecific, generat-
ing a large percentage of false positives [BSC+ 02]. In other words, theoretical
prediction of promoters is still one of the greatest challenges in bioinformatics.
Bioinformatics is a recent field of study, and involves the use of computers
and algorithmic knowledge to solve biological problems. In the last 30 years,
thanks to the huge improvements in computers and databases, biologists have
started to use them to cross biologic data with the hope of finding new informa-
tion. Later on, with the introduction of multidisciplinary researchers, the data
was converted to information structures and now almost every biological field
can use computers to enhance their work.
On the interface part, software tools for bioinformatics range from simple
command-line tools, to more complex graphical programs and standalone web-
services. With the increasing availability of Internet tools and the invention of
1 The reader might find in this introduction some biological terms. All of them are explained

in chapter 2.

2
remote data exchanging protocols, it is now possible to run software in remote
servers, obtain the results, parse them, and send the results to another machine
which will display them in a graphical window.
Databases are huge and scattered, and there is room for improvement in data
display interfaces as well as new methodologies to extract more information from
that data. The trend is moving from using only one algorithm and one interface
—a terminal running a script— to the use multiple sources of information and
multiple interfaces.

1.1 Definition of the problem and objectives


Despite the recent sequencing of the human genome2 [VAM+ 01], biologists
know little about the mechanisms regulating the genes. One of the most impor-
tant regulators are promoters, regions of DNA located near the genes. Specific
proteins recognize these promoter areas and start the transcription of the gene.
Promoters can be discovered by experimental means, in vitro, but these
methods are slow and difficult. Since the sequencing of the human genome,
early bioinformatic tools have been trying to discover new promoters [DGZ01],
changing the computational approach with every new genetic discovery.
The first algorithms [OcLNR02] relied on the DNA sequence itself, while the
most recent techniques [ASRdP08] use structural profiles of the DNA as the
base for their predictions. In this Thesis, we will follow the most recent trends
on promoter detection, designing a platform to calculate structural profiles from
a DNA sequence, and then a promoter detection software which will use these
profiles to predict promoter location.
In the first part, we will implement DNAlive, a multi-interface platform to
compute up to 29 physical descriptors of a DNA molecule (listed in appendix.
B). For the second part, following the community-standard EGASP [BBB+ 06]
experiment, we will analyze the DNA profiles of a public DNA database and use
Neural Networks to improve the prediction power of a state-of-the-art software
developed by the MMB group, ProStar.

1.2 Application of the AI


The first part of the Thesis, the construction of DNAlive, is mainly a computer
engineering problem. It consists on the development of two interfaces to allow
the remote execution of DNA descriptors. The interfaces are a web page and
bioinformatic web services.
To implement the descriptors, two basic algorithms were used: geometric
operations on the input data and fuzzy logic, where the resulting scalar value
of an operation could be modified by conditional distributions. However, the
methodology for each one is clearly specified on their respective paper (see
appendix B). Thus, even in the cases where some AI technique is used, it was
not developed during this thesis, and as such it is not explained here.
On the second part, ProStar 2 makes extensive use of AI methodologies. The
core techniques that will be used to improve the previous version, and detailed
in ch. 5, are:
2 All terms in bold are included in an Index at the end of the document.

3
• Analyze the DNA descriptors and determine which are more valuable to
detect promoters, using data mining techniques.
• Promoters will be classified in four groups, each one representing a dif-
ferent promoter type. Then, each one will be treated separately to gain
specificity.
• A neural network for each group will be trained and then used to predict
the presence of a promoter in a new DNA sequence.

The promoter detection task has usually been labeled as a symbolic problem.
The DNA sequence in its naked form was the only information used, and each
element of the sequence is a symbol (A, C, G, T). Classic techniques include
pattern matching or stemming, or even Hopfield networks to reduce noise.
By using DNA numerical descriptors instead of the plain sequence, the prob-
lem turns to be subsymbolic. After computing the descriptor, the original se-
quence is discarded, only the numerical values are kept. The use of this method
allows to plot a graph of the descriptor signal along the sequence, and some
important DNA elements can be discovered by looking at the signal. Unfortu-
nately, this is not yet the case for promoters.
The descriptors analysis will help to reduce the search space and input noise.
These values have biological meaning, but it is unknown which one of them best
describes a promoter. Using a comparison, looking for promoters by using the
DNA temperature value could be like trying to define a musical composition by
the color of the speaker it plays on.
Statistical methods will be used to compute the relevance of each descriptor,
starting with a Pearson correlation and then running a Principal Components
Analysis. Correlations will tell which descriptors are redundant, and the PCA
will determine the weight (i.e. relevance) of each descriptor.
Promoters can be clustered in different ways, based on some of their proper-
ties. In this thesis, they will be classified by analyzing the presence or absence
of the two most common elements, forming four groups. The classification is
quite simple, and one interesting conclusion will be whether these groups are
differentiated by physical descriptors or not.
In order to get a class predictor (promoter or non-promoter), many classifiers
could be used. For this problem, neural networks were chosen, as it is a very
flexible method, specially suited to treat sets of inputs as signals and to build
models that are, to some extent, noise-resistant.
Other alternatives, such as support vector machines, were discarded because
the lack of a kernel function. Self organizing maps are used in [ASRdP08] but, as
we are going to cluster the promoter instances beforehand, it makes no difference
compared to a neural network.
Other AI techniques were discarded because the lack of application. Some
types of learning, like reinforcement learning, cannot be applied, as the answer
space is not big enough. In a same manner, it is impossible to use deductive
learning or rule-based learning, because there is little knowledge about the so-
lution, no model has been found to define a promoter, and thus the model has
to be created.
This thesis is a clear example of the application of the AI in many research
fields. In bioinformatics, promoter prediction is a main problem which will lead
to better genetic research.

4
The only way to validate the predictions is to compare them to an annotated
DNA database, but these databases are very small. When a predictor performs
well, research groups use it to find places where it detects unknown promoters
with the most confidence value, and then look for them in vitro, in a biology
laboratory.
Not only is it important to discover promoters, but also improve the confi-
dence values and reduce the number of false positives. This thesis’ results will
provide light on the application of new methodologies to a current problem.

1.3 Structure of this thesis


This document is organized as follows. After the introduction, in chapter 2 we
introduce the biological concepts behind this thesis. In chapter 3, current bioin-
formatics software will be presented, plus a series of gene predictors, including
the first version of ProStar.
Chapters 4 and 5 explain the methodology that we have used to implement
the DNAlive platform and ProStar 2 promoter predictor, respectively. After
presenting the methods, chapter 6 shows the results for both projects, and
chapter 7 analyzes these results, comparing them to the initial objectives.
Two appendices are included. Appendix A contains the published paper for
DNAlive, and appendix B includes a comprehensive description of the DNA
physical properties included in DNAlive, with a reference to their original pub-
lications.

5
Chapter 2

Biology

DNA was first isolated by the Swiss physician Friedrich Miescher who, in 1869,
discovered a microscopic substance in the pus of discarded surgical bandages.
Since then, much effort has been put in the investigation of this element and its
role in the living beings. It was not until 1953 when, based on X-ray diffraction
images taken by Rosalind Franklin and the information that the bases were
paired, James D. Watson and Francis Crick suggested what is now accepted as
the first accurate model of DNA structure in the journal Nature [CW53].
This chapter defines the terms that will be used in the rest of the Thesis,
and provides a basic but detailed introduction to the concepts that are needed
to understand this work.

2.1 Central dogma of the molecular biology


Living organisms can be divided between prokaryotes and eukaryotes.
Prokaryote cells lack a nucleus, cell membrane, and many other structures
that are seen in eukaryotic cells; in general, they are less evolved and simpler.
Eukaryotic cells are present in higher animals and their structure is very com-
plex; a comparison of both cells is pictured in fig. 2.1.
Prokaryotes also contain genetic material —DNA—, but it is not packed
(more on this in section 2.5) and its behavior is slightly different than eukaryotes.

Figure 2.1: Simplified models for the cells of eukaryotes and prokaryotes.

6
Figure 2.2: A chunk of DNA in its stable 3D form (Source: Wikipedia)

For this Thesis, unless specified, all references to cells are meant to be for
eukaryotes, the ones present in human organisms. As stated, every cell of the
human body contains a nucleus, where the DNA is surrounded by a liquid
composed of water and other molecules.
It is well known that the DNA is a double helix (see fig. 2.2) that contains
genetic information. However, this big molecule has very intricate physical and
chemical structure, and plays a critical role in genetics.
An easy metaphor to understand the DNA structure is to view it as a long
string, composed of character sequences. Each of the four letters that form the
DNA (A, C, G, T) represent a small molecule that has a name, with its atoms,
its bonds and its physical forces. These letters are named Adenine, Cytosine,
Guanine and Thymine, and their generic name is base. Each one has a sugar
molecule which acts as a scaffold, the backbone. The combination of a base
with a sugar forms a nucleotide1 .
A chunk of DNA expressed as GGCAATTACGACGGTATAACT means that there
are two Guanine molecules, followed by a Cytosine, followed by two Adenines,
two Thymines, and so on.
Furthermore, every nucleotide is facing another nucleotide, forming the afore-
mentioned double helix. The nucleotides always pair themselves in the following
fashion: C pairs with G, while A pairs with T (or U, in RNA), and vice versa. For
example, AACT always faces TTGA.
By convention, one of the strands is read from the top to the bottom, while
the other is meant to be read from the bottom to the top. The “top” end is
called the 5’ end, and the “bottom” is called 3’ end. A picture of the chemical
representation of the DNA can be seen in figure 2.3.
Today, it is known that the DNA encodes the information that expresses all
the human body. It has information on the hair color, the gender, the number
1 However, in this thesis, base and nucleotide are used as synonyms.

7
Figure 2.3: The DNA chemical structure. The left strand is read from top (5’
end) to bottom (3’ end), while the right strand is considered to be inverted. As
presented here, each nucleotide is composed of different atoms, and the structure
is held by atomic bonds. The “backbone” is the scaffold which subjects the
nucleotides. (Source: Wikipedia)

of toes per feet, and so on. Many things about the DNA have already been
discovered, like the translation process, but some others still remain unknown,
like the gene regulation.
The central dogma of the molecular biology [Cri70] enunciates the
normal flow of information that regulates genetics on living beings. Basically, it
states that the DNA holds the genetic information, and the different means of
transferring it. This flow can be interpreted as a graph, as depicted in Figure
2.4.
Each DNA sequence encodes some function in the body. One of the most
important, however, is the production of proteins. Proteins are very complex
molecules that, in turn, are formed by smaller molecules called amino acids.
In every three DNA nucleotides the code for a protein molecule is expressed; for
example, nine DNA nucleotides are equivalent to three amino acids.
In fact, there is another molecule involved in this process, the RNA, which
is very similar to DNA but usually appears in a single strand form. It acts
as a mold, being the intermediary between the DNA chain and the protein.
Furthermore, in RNA, the nucleotide uracil (U) substitutes the DNA’s thymine.
The DN A → P rotein process consists of two phases:

• Transcription, where the DNA generates a piece of RNA.


• Translation, when the above RNA produces a protein. That protein then
will express a gene.

The Central Dogma also explains a third information transfer: the DNA
replication, used when the DNA duplicates, for example, when transferring

8
Figure 2.4: Central dogma of the molecular biology. The blue arrows determine
the normal information flow, while the red arrows point to special cases. (Source:
Wikipedia)

information to the progeny. In this three-way game (DNA, RNA, Proteins)


there might also be other ways of transferring information, but they are very
rare.
Even though the DNA/RNA express the Proteins, these latter molecules are
essential parts of organisms and participate in every process within cells. They
process food, build up the cells, and they even generate other DNA sequences.
For example, the process of DNA duplication is carried out by a complex group
of proteins that unwind the helix and, using another protein named DNA
Polymerase , copy or replicate the master template itself. The proteins which
regulate gene expression, binding to the DNA and activating the transcription
process, are called transcription factors.

2.2 Genes
Genetics is the area of biology that studies genes, which is the basic unit of
DNA information. All these are, in fact, DNA sequences placed strategically
in the body cells. Using naive examples, one can say that, if the sequence
CCTTACAAAATAGGGTG is present at about the 12,413,574th position of someone’s
21st Chromosome, then that person is going to have green eyes.
In the last paragraph there were three remarkable pieces of information:
• the DNA sequence,
• the position and the Chromosome where it is placed (the Transcription
Start Site, or TSS), and

9
• its function.
Today, many of the human genes rest still undiscovered, mostly because one
—or more— of the above information is unknown. The most usual case is that
scientists know that between the 12,400,000th and the 12,450,000th position
of the human’s 21st chromosome there is some sequence that regulates some
unknown function, nothing else.
Nowadays there are very powerful computers that are capable to handle large
character sequences and work with them. The whole Human Genome lengths
about 3 Gigabases, three thousand million letters. But, how can we extract any
data from just plain letters?
Given the fact that the DNA has physical properties that define its flexibility,
and that a protein is not a language parser that reads the whole 3 Gigabases
before attaching somewhere, the conclusion is clear: all interactions that happen
in the human body and, in extension, everything regarding DNA and proteins,
are ruled only by physical forces (e.g. electro-chemical forces).
Different families of genes are expressed by different proteins. For instance,
Polymerase II (Pol-II), a very large protein with more than 10 subunits, is
the main player in the transcription of genes encoding for messenger RNA
(mRNA), the type of RNA that transports the information from the DNA to the
proteins. However, it is also clear that Pol-II is not able to start transcription
by itself, but needs of a large number of additional proteins.
Those proteins, named transcription factors (TF), create a large protein
cluster bound to DNA, the pre-initiation complex, which precedes the tran-
scription of the corresponding gene [SK03]. The region 200-300 bp2 upstream
the core promoter (see sec. 2.3), where the TF binds, defines the proximal
promoter area, where multiple transcription factor binding sites are located.
Additional signals are received from enhancers which can be bound far —even
thousands of bases away— in the sequence of DNA, but that the DNA structure
probably locates close enough as to allow interaction with the pre-initiation
complex, the proteins involved in gene translation.
Genes can be divided in two big groups: coding and non-coding. The
coding sequence determines what the gene produces, while the non-coding se-
quences are known to be genes, but somehow they do not seem to be translated.
Coding genes account for about 30% of the total, while non-coding account for
70% of all genes.

2.3 Promoters
A promoter is a small region (hundreds of bases) located just before a gene,
and it is the place where the transcription factors attach to start expressing a
gene.
They are of extreme relevance because of several reasons. First of all, they
are strictly related to genes, and finding a new promoter on a relatively deserted
DNA sequence usually leads to find a new gene. Second, promoters regulate gene
expression, which looks like Figure 2.5.
The core promoter is the region immediately upstream of the TSS, where
the transcription initiation complex assembles. It is located upstream of the
2 Base Pairs, a size unit meaning a base A, C, G, T and its complementary T, G, C, A

10
Figure 2.5: The process of gene expression. A Transcription Factor (blue, thin
and curly) has bound to the DNA (orange, double strand) and is producing
the RNA (green, in between the top opened DNA strands) which will, in turn,
express the gene in the final Protein. (Source: Wikipedia)

coding part of the gene, sometimes up to several thousand base pairs, and is
responsible for basal transcription as well as transcriptional regulation of the
gene it is linked with.
Biologically speaking, there are many signals that help us detect promoters
[AD07]. One is the Initiator Element (Inr), located around the TSS and
can work independently or synergistically with the TATA box. However, the
prevalence of the Inr element in mammalians is not clear, but it seems to be
quite abundant in Drosophila3 . TATA boxes and CpG islands are very clear
signals that have been studied thoroughly, and they will be reviewed.
Still, the presence of a TATA box or a CpG island is not a requirement for
the presence of a promoter, since there exist promoters without them. Similarly,
their absence does not imply that there is no promoter.

2.3.1 TATA boxes


Early analysis of common eukaryotic signals at proximal 5’ (see 2.3) upstream
region of known genes revealed the presence of some over-represented motifs
which were demonstrated to serve as signals for placement of the pre-initiation
complex. The TATA box is the strongest of these signals and is recognized
by the key TF IID complex. It has a consensus sequence TATAAAA, but large
3 Drosophila is a species of fruit flies, used in laboratories because they breed fast, DNA

recombination and mutations are common, and they share a lot of DNA with humans.

11
deviations from this consensus have been found in different genes [SK03].
It is generally found in the 25-30 upstream of the transcription start site
(TSS), but again this distance can change depending on the organism [HS89].
Traditionally, the TATA box was supposed to be present in around 30% of
genes, but it is not present in oncogenes, growth factors and house-keeping
genes [SS03]. Furthermore, recent experiments performed within the ENCODE
project4 strongly suggest that the number ot TSS linked to TATA-box can be
even lower [BTC+ 06].

2.3.2 CpG islands


DNA methylation is a type of chemical modification of DNA without changing
the original sequence. It involves the addition of a methyl group (CH3 ) to the
DNA with the effect of reducing gene expression. DNA methylation at the 5th
position of Cytosine has been found in every vertebrate examined [GGF87].
DNA methylation may impact the transcription of genes in two ways. First,
the methylation of DNA may itself physically impede the binding of transcrip-
tional proteins to the gene and secondly, and likely more importantly, methy-
lated DNA may be bound by proteins that form compact, inactive chromatin.
CpG islands are unmethylated long (500-2000 bp) segments of DNA, with
with at least 50% C+G content, and the number of CpG dinucleotides being at
least 60% of what could be expected from the C+G content of the segment. The
algorithm is further explained in section 5.2.1.
Promoters associated to CpG islands have multiple TSS that span a region
of 100 bp or more [DGZ01] and usually lack other signals like TATA boxes, DPE
or Inr elements [STS+ 01]. CpG island-associated promoters seem to be rapidly
evolving in mammals, whereas TATA box promoters are more constrained.
Between 60-90% of all CpGs are methylated in mammals, and unmethylated
CpGs are grouped in CpG islands. In many disease processes such as cancer,
gene promoter CpG islands acquire abnormal hypermethylation, which results
in heritable transcriptional silencing, avoiding the expression of some genes.

2.4 Physical properties of the DNA


The mechanical properties of DNA5 are directly related to its structure.
Every process which binds or reads DNA is able to use or modify the mechanical
properties of DNA for purposes of recognition, packaging and modification. The
extreme length, relative rigidity and helical structure of DNA has led to the
evolution of techniques to compact a cell’s DNA. The properties of DNA are
closely related to its molecular structure and sequence, particularly the weakness
of the hydrogen bonds and electronic interactions that hold strands of DNA
together compared to the strength of the bonds within each strand.
One of the most important physical property of the DNA is its own shape.
DNA appears in three conformations: A-DNA, B-DNA and Z-DNA, as
4 ENCODE (the ENCyclopedia Of DNA Elements) is a public research consortium

launched by the US National Human Genome Research Institute in September 2003. The
goal is to find all functional elements in the human genome, one of the most critical projects
after the successful completion of the Human Genome Project.
5 In this document, the terms DNA physical property, DNA descriptor and DNA mechanical

properties are used as synonyms.

12
depicted in figure 2.6. B-DNA is the standard form, and the B name is used
only for description purposes, as it is commonly referred just as “DNA”. A-DNA
and Z-DNA differ significantly in their geometry and dimensions to B-DNA,
although they still form helical structures.
The A form is likely to occur only in dehydrated samples of DNA, such as
those used in crystallographic experiments, and possibly in hybrid pairings of
DNA and RNA strands. It is a right-handed double helix fairly similar to the
more common and well-known B-DNA form, but with a shorter, more compact
helical structure.
Segments of DNA that cells have methylated (see 2.3.2) for regulatory pur-
poses may adopt the Z geometry. It is a left-handed double helical structure in
which the double helix winds to the left in a zig-zag pattern, instead of to the
right, like the more common B-DNA form. There is also evidence of protein-
DNA complexes forming Z-DNA structures.
Other conformations are possible; (C)ovalent mitomycin-DNA [SFL+ 95],
(D)elta-DNA [SMHK01], (E)ccentric-DNA [VEH00], (L)ambda-DNA [SCHH82],
(P)auling-DNA [ABLC98] and S-DNA [CLH+ 96] have been described so far, al-
though they only appear in some specific organisms and are not common.

Figure 2.6: The structures of A-DNA, B-DNA and Z-DNA. B-DNA is the canon-
ical form, discovered by Watson and Crick 50 years ago. A-DNA is typical on
RNA, but also present in DNA sometimes. Notice how the Z-DNA winds to the
left, whereas the A- and B-DNA forms wind to the right. (Source: Wikipedia)

At another structural level, we can also find DNA triplexes and quadruplexes.
These DNA structures are composed of three or even four strands, instead of
the typical two-stranded model.
The ability of the DNA to form these structures is known since the 50s-
60s, but in the last years many studies have brought back the interest for them
[SCH96], due to their possible implication in biological processes such as the
transcription and to their potential biotechnological applications, like a viable
information-encoding system. [LGK05]. Expanded DNA, especially quadru-

13
Figure 2.7: A triplex structure, side and top images. The two original strands
are depicted in red, and the extra strand appears in green.

plexes, encode four bases of sequence information, and it forms antiparallel


double helices of high stability and (generally) high selectivity.
A triple helix (fig. 2.7) is formed when a third string (RNA, DNA or a
combination of both) is placed in the big hole of the double DNA helix and their
bases interact with the hydrogen bonds. The triplex bases always form (CG)C
or (TA)T triads.
The quad-helix (fig. 2.8, usually named quadruplexes or G-DNA, is an
unusual DNA structure that is found at the bottom edges of the chromosomes,
named telomeres. These regions are very rich in guanines, and are very im-
portant in the DNA replication because they allow the binding of the comple-
mentary strand. They are mostly present in the telomeric regions because they
help to preserve and compact the DNA.

Figure 2.8: A quadruplex structure, side and top images. The whole structure
is stabilized by a N a+ or K + ion in the center.

14
Figure 2.9: A nucleosome (green) bound to the DNA (red to blue scale).

2.5 Nucleosomes and Chromosomes


Human DNA is a very long string, about 3 billion bases in humans, divided into
24 chromosomes. As each base is 0.33 nm long, that makes about 1 meter of
DNA in each human’s cells; that is why DNA forms higher order structures.
The DNA is compacted many times with the help of proteins. The most
basic one is the Nucleosome, a big structure that glues 146 base pairs in a
loop shape, as can be seen in figure 2.9.
The DNA folding does not stop with nucleosomes. The resulting structure
is compacted again and again, with the help of many different proteins, forming
a substance named chromatin (literally: colored material, as it was discovered
by staining the nucleus of the cells). The structural entity of the chromatin is
the chromosome. The different compaction levels are depicted in fig. 2.10.

2.6 Concluding remarks


This chapter has presented a quick reference with all the definitions that are
required to understand the biological background behind this Thesis. Further
chapters reference some of this sections when needed.
The most relevant sections of this chapter are Promoters (2.3) and Physical
properties(2.4). They define the basis for ProStar2 (ch. 5) and DNAlive (ch.

15
Figure 2.10: This picture shows how the DNA is compacted many times, forming
chromosomes. The resulting complex, composed of DNA and proteins, is called
chromatin. (Source: Wikipedia)

4), respectively, and more on these concepts will be explained in those chapters.
Next, we will overview some of the available tools which apply these concepts
in bioinformatics software.

16
Chapter 3

State of the art

Bioinformatics covers a wide range of topics, which could be naively divided


into databases and algorithms. Databases are especially extended and have be-
come the most popular resource, but here only the most relevant are presented.
Regarding DNA manipulation, the most accessed ones are Genome Browsers
and BLAST-like software. Genome browsers allow the visualization of data in
an intuitive manner, allowing the integration between databases. BLAST is a
sequence searcher, and it has different versions for DNA, proteins and other
structures.
Promoter prediction software is also abundant [BBB+ 06], however only re-
cent methods have reached an acceptable level of accuracy. In this section two
programs are analyzed, one being ProStar, and the other EP3, a recent work
which shares some similarities with ProStar 2.

3.1 Genome browsers


As the genomic data grows faster and faster, researchers need tools to visualize
the DNA sequences and their annotations. There is plenty of software aimed at
displaying plain DNA structures, but the user needs to integrate further data
by hand.
In 2002, a year after the human genome was fully sequenced, Kent et al.
[KSF+ 02] developed the UCSC Genome Browser1 . Its main objective is
to provide a rapid and reliable display of any requested portion of the genome
at any scale (fig. 3.1), together with several dozen aligned annotation tracks
(databases). The browser also displays a huge amount of data related to the
DNA, like RNAs, known genes, cross-species data and so on. Furthermore, users
can add their own data tracks and display them along with UCSC’s ones (see
fig. 3.2).
Recently, the UCSC team has published another article with an update on
the project [KKBB08]. The project is still alive, and there have been great
improvements and many spin-off projects, like a genomic wiki.
1 Publicly available at http://genome.ucsc.edu

17
Figure 3.1: The UCSC Genome Browser plotting many DNAlive descriptors.
On the top, there are the genomic coordinates; on the left, the name of each
track.

Figure 3.2: Managing tracks in the UCSC Genome Browser. The interface
provides many alternatives to display the data, allowing a full customization of
each track.

18
3.2 Sequence search
BLAST (Basic Local Alignment Search Tool) is an algorithm for comparing
primary biological sequence information, such as the aminoacid sequences of
different proteins or the nucleotides of DNA sequences. A BLAST search enables
a researcher to compare a query sequence with a library or database of sequences,
and identify library sequences that resemble the query sequence. For example,
following the discovery of a previously unknown gene in the mouse, a scientist
will typically perform a BLAST search of the human genome to see if humans
carry a similar gene.
BLAT is a BLAST-like tool designed to quickly find DNA sequences of 95%
and greater similarity of length 25 bases or more. It keeps only an index of the
entire genome in memory, which takes up a bit less than a gigabyte of RAM. The
genome itself is not kept in memory, allowing BLAT to deliver high performance
on a reasonably priced box. Then, a user can input a DNA sequence, and the
BLAT server will quickly (5-6 seconds) return the genomic coordinates where
that sequence is located; that is very useful to display annotated DNA from a
sequence which was thought to be unknown.

3.3 DNA dynamics


Molecular dynamics (MD) is a form of computer simulation in which atoms
and molecules are allowed to interact for some of time under the laws of physics,
allowing the view of the motion of the atoms. Molecular systems are composed
of a huge number of elements, so it is impossible to find the properties of such
complex systems analytically. MD simulation uses numerical methods to avoid
this problem. Its laws and theories stem from mathematics, physics, and chem-
istry, and it employs algorithms from computer science and information theory.
Today it is applied mostly in materials science and biomolecular simulation.
MD simulations require a lot of computational power. Nowadays, simulating
the movement of an average protein (10,000 atoms) for 1,000 nanoseconds2 in
128 CPUs of a machine like the Mare Nostrum can take as much as 6 months of
time and occupy 10 TB of data. Most MD, however, are simulated for 20-100
ns, reducing these values to a more handleable time. Of course, after running
the simulation, the results must be extracted and analyzed, which accounts for
most of the time of the researcher.
The general algorithm ruling MD is simple to understand, but very difficult
to implement. Basically, it iterates the following process through time (typically,
10−15 seconds):

• Get the forces for each atom and its acceleration


• Move each atom to its new position

To get the atomic forces, a very CPU-intensive operation is used to compute


the force field or energy potential. Then, corrections need to be applied, to
fix the natural vibrations of the atoms. In addition, most MD use an explicit
solvent, that is, the atoms need to be surrounded by water molecules, which
exponentially increases all the calculations.
2 MD time is measured in nanoseconds

19
Because of the complexity and expensiveness of classic MD, a simulation of
the simulation technique is performed to determine basic molecular dynamics
with a Monte Carlo (MC) method.
The MC approach relies on the distribution of random numbers to approx-
imate expensive physical or mathematical calculations. Instead of trying to
reproduce the dynamics of a system, it generates states according to appro-
priate Boltzmann probabilities. It employs a Markov chain procedure in order
to determine a new state for a system from a previous one. According to its
stochastic nature, this new state is accepted at random. The avoidance of
dynamics —in Markov processes, each state is independent from the previous
one— restricts the method to studies of static quantities only, but the freedom
to choose moves makes the method very flexible.

3.4 Gene predictors


NSCAN, a modified version of TWINSCAN [GB06], is a paradigm of pro-
gram detecting promoters based on gene structure. The program uses gener-
alized Hidden Markov Models to find normal signals of genes locating the
5’ end and conserved regions upstream, guessing then the promoter position.
However, most of the programs for promoter detection are based on the idea
that there are subtle sequence signals associated to promoters.
Thus, many algorithms are trained to recognize basic signals like the TATA
box and/or Inr, CpG islands or regions with a large population of targets for
Transcription Factors. The location of CpG islands has focused special effort
and several methods have been independently trained to locate CpG-positive
and CpG-negative promoters. Several of these methods use compositional rules,
at the trimer, pentamer or hexamer levels3 , or in most sophisticated versions
Hidden or Iterative Markov chains trained against known promoters. Finally,
some methods like FirstEF [Dav03] or Dragon Promoter Finder [BSC+ 03],
Dragon Gene Start Finder [BS03] or promH [SS03] take advantage of the
predicted gene structure to help their sequence-trained models to locate pro-
moters.
The diffuse character of sequence signals at promoters indicates that factors
other than the specific hydrogen-bond interaction between nucleotides and bind-
ing proteins modulate the recognition of target DNA fragments. As suggested
by others [KB05], one of these additional factors can be the physical properties
of DNA, which can control the degree of accessibility of the target sequences
to binding proteins through the modulation of chromatin structure, the trans-
mission of enhancers or proximal promoter information to the core promoter
region, or the formation of protein aggregates in the pre-initiation complex.
The fact that DNA at promoter regions displays different physical properties
than the rest, specially near the TSS, has been clearly probed by in prokary-
otes and with less clarity also in different eukaryotes [PBCB98]; [ONGR01];
[FSD+ 05].
3 i.e. grouping the bases in sets of three, five or six elements

20
3.5 ProStar 1
After this exhaustive review of the promoter detection techniques, the MMB
group decided to implement a mechanism to derive all the known physical prop-
erties and then analyze possible differential properties of DNA in regulatory
regions in vertebrates. As expected, parameters are often correlated and the
informative content is not always easy to determine [OcLNR02]. Beyond the
idea that these parameters with stronger signals near TSS are supposed to play
a determinant role in regulatory mechanism, they searched for this reduced set
of parameters that were supposed to be able to cooperatively predict a promoter
region.
In ProStar’s case, first of all, the six helical force constants were calcu-
lated by means of MD simulations [PMSS07], as depicted in figure 3.3. Four
duplexes containing several replicas of every possible combination of two nu-
cleotides were used, plus four extra sequences related to CpG islands and TATA
boxes: GCCTATAAACGCCTATAA, CTAGGTGGATGACTCATT, CACGGAACCGGTTCCGTG and
GGCGCGCACCACGCGCGG. The trajectories were manipulated to represent the de-
formability of a given step along the three rotations and the three translations.
The dataset was gathered from protein-coding genes annotated by the HA-
VANA group, following the EGASP workshop rules[BBB+ 06]. Then, the method
was trained for promoter recognition with a collection of 500-nts4 sequences
that comprised intervals of 250 nucleotides upstream and 250 downstream of
the training TSS set. For the negative set, random 500-nts sequences from
transcribed regions were selected, making sure that they did not overlap with
the positive set. For the recognition of the strand, the method was trained with
a collection of DNA sequences that comprised 900-nts upstream and 900 down-
stream of the same TSS. The reverse complementary sequences were taken as
the negative set.
Using the MD derived parameters, any DNA sequence of size n can be de-
scribed as a 6-dimensional deformation vector
v = (twist, tilt, roll, shif t, slide, rise). For a given deformation, the values asso-
ciated to every dinuecleotide step in the sequence are summed and then divided
by n − 1. For example, the twist deformation score of sequence ACGC would be
(0.036[AC]+0.014[CG]+0.025[GC])/3 = 0.025. The 6-dimensional vector of the
same sequence would then v(ACGT ) = (0.025, 0.033, 0.022, 1.200, 2.547, 8.230)
(see fig. 3.4)
To classify the input sequences into the promoter class (kx ) or the non-
promoter class (ky ), the Mahalanobis distance is used. Computing the physical
properties of every sequence of the dataset a set of vectors for every class is
concluded (X for class kx and Y for ky ). The Mahalanobis distance DM
between the set of vectors X and Y is defined in eq. 3.1:

DM (X, Y ) = (µx − µy )t C −1 (µx − µy ) (3.1)


Where µx and µy are the average vectors of the sets X and Y and C −1 is the
covariance matrix of X ∪ Y . The decision function g of a specific 500-nts DNA
sequence with a description vector s to a class ki with i =< x, y > is defined in
equation 3.2.
4 Read as “500 nucleotides long”

21
Every possible
combination of
CTGA

MD simulations

6 helical
force
constants

Figure 3.3: ProStar1: Calculus of the force constants

Negative set:
Positive set:
Random
Coding genes
sequences from
from HAVANA
HAVANA

6 helical Calculate the


force force constants
constants

Promoter
database

Figure 3.4: ProStar1: Generation of the promoter database

22
6 helical
User input
force
constants

Calculate the Promoter


force constants database

Mahalanobis
distance classifier

Promoter, No
strand Promoter

Figure 3.5: ProStar1: System architecture

g(s, ki ) = wkt i s + wki ,0 (3.2)


−1
Where wki = C µi and wki ,0 = −0.5µti C −1 µi .
When g(s, kx ) > g(s, ky ),
the sequence is classified as a promoter. Even so, the confidence of the decision
can be modulated according to a normalized score defined in equation 3.3. If the
score is greater than a specific threshold (set +1 by default), then the sequence
is flagged as a promoter.

g(s, kx ) − g(s, hy )
score(s) = (3.3)
g(µx , kx ) − g(µx , ky )
The strand is predicted by recognizing the upstream/downstream signal
asymmetry using a statistically discriminator based, again, on Mahalanobis met-
rics. Finally, very close clusters of predictions (using a 1000-nts window) are
unified in a single hit. A graphical view of the system is depicted in fig. 3.5.
ProStar’s method is conceptually and computationally simpler than any
other general promoter prediction algorithm as it does not require any addi-
tional information, such as conservation of gene structure across species, pres-
ence of CpG islands, TATA-boxes, Inr elements or any other sequence specific
signals. Due to its simplicity ProStar can be in principle applied even in cases
where promoters are located in unusual genomic positions.

23
Figure 3.6: Physical properties study for EP3

The results of these comparisons show that despite its simplicity, ProStar
behaved better than most of the other methods, similar to two algorithms that
use gene structure for prediction (fpom and firstef ) and only nscan, that is based
also in multi-species homology, provides more accurate results for the reference
set of genes. (see [GPTO07], pages 5–8)

3.6 EP3
Abeel et al. developed EP3 [ASB+ 08], a novel approach for predicting pro-
moters in whole-genome sequences by using large-scale structural properties of
DNA. EP3 requires no training, is applicable to many eukaryotic genomes, and
performs extremely well in comparison with the best available promoter predic-
tion programs to the date of publishing. Moreover, it is fast, simple in design,
has no size constraints, and the results are easily interpretable. Their method
also has been tested on 12 additional eukaryotic genomes, including vertebrates,
invertebrates, plants, fungi, and protists.
EP3 is very relevant for this Thesis because they analyze the relationship
between different physical properties and the known promoter groups. Specif-
ically, the variation of RNAP II is plot for each promoter descriptor: TATA
boxes, CpG islands, TFIIB and Inr (see sec. 2.3). This graphic is depicted in
figure 3.6
While EP3 does not outperform its peers by much, the program has sev-
eral additional advantages compared with other promoter predictors. EP3 re-
quires no training or parameter tuning, unlike other programs that need exten-
sive amounts of experimentally determined data for the training of their model
([OSHN00]; [SKW00]; [DGZ01]; [DH02]; [BSC+ 03]). When working on a ge-
nomic scale, speed and memory requirements also are of importance. EP3 is
very fast (for instance, it takes less than 1 h to annotate the complete human

24
genome), requires little memory, and can thus be run on a home computer; in
contrast, some programs require a computer cluster of 80 machines for nearly
a week to process the human genome. Besides performing very well, especially
in light of its simplicity, EP3 can handle many eukaryotic genomes without
modifications.

3.7 Concluding remarks


In this chapter we have described the most relevant works for gene prediction
that currently exist. ProStar 1 is somehow the base for this thesis. EP3, by
Abeel et al. [ASB+ 08], is relevant because it was published the same week than
ProStar, and using a very similar methodology gets similar results, although
for some specific promoter groups the accuracy varies between both programs.
Also, its authors performed a study of the relationship between promoter signals
and physical properties.

25
Chapter 4

Creating a platform for the


analysis of DNA: DNAlive

In this chapter, the whole process for constructing DNAlive will be presented.
A typical software engineering process was used since the beginning, composed
of the following steps:

1. Define an objective
2. Gather the requirements

3. Design the architecture


4. Design the system
5. Implement the software
6. Test the software

4.1 Objective and requirements


The main objective of DNAlive is to integrate in one public platform all the
descriptive information on the human genome with the physical properties of the
DNA. In section 3.1 we presented the UCSC Genome Browser, the most popular
genomic annotation tool, and we wanted to use it to display the information
generated by the physical properties predictors.
Furthermore, we also decided to implement a 3D and 4D (animated 3D)
viewer, because 2D maps sometimes are not visual enough; also, the platform
needed the ability to search in databases for proteins which bind to a DNA and
physically compute this binding to the double helix.
In general, the software needs to be able to to run many prediction and
structural algorithms for a given DNA sequence, everything automatic, so that
the user can get many data without being required to input further information.
Data integration and calculus automation are the two main pillars of DNAlive.
The system should also have two main interfaces: a web page and web
services. The first one should be compatible with the most used version of

26
every major internet browser. The web services must be compatible with other
services and exchange data seamlessly.
The web page must be easy to use, work in real time and provide enough
feedback so that the user always knows what to do next. For that, the perfor-
mance should be tweaked to the maximum for all the data to be calculated in
a few seconds. When the system finds that some calculus is going to take more
than a minute, the user should be warned and asked for confirmation1 .
Regarding reliability, even though we will ensure that no data is lost, it is not
a main requirement, as all the data can be regenerated and all the steps can be
reproduced in a few seconds. Security is also important for the server part, and
we must ensure that the software has no exploits. For the client part, however,
no critical or confidential data is handled, so no extra security measures (SSL,
passwords) need to be taken.
Finally, one of the most important requirements was modularity. As we
needed to construct two interfaces for accessing DNAlive, every operation must
be modularized.

4.2 System architecture and design


With modularity in mind, we generated a list with all the operations that we
needed to implement and we built up an architecture diagram (figure 4.1)
The user only needs to input genomic coordinates or a DNA sequence. They
are interchangeable in most cases, as the software can use BLAT [Ken02] to
look for the coordinates for a given sequence or retrieve a chunk of DNA for a
specific genome.
Once the input data has been processed, the user can calculate the physical
properties for the sequence and plot a 2D map of them. If these coordinates
belong to a genome, the UCSC genome browser can be used for the plot. Else,
we need to implement our own 2D plotter.
Then, we generate the DNA 3D structure from the sequence, calculate its
physical properties, and plot the 3D or 4D images. Bound proteins can also
be searched for, as an optional step. We will run an algorithm that looks for
proteins on a database given a DNA sequence, and will return the results. After
that, another algorithm will try to bind the proteins to the DNA sequence, and
then return a DNA-Protein structure with the best matches.
It should be noted that, for the 4D plot, no physical properties are com-
puted. This was a design decision, derived from software constraints. When we
tested the animation software we discovered that painting physical properties
on a 4D animation was very resource-demanding and the system went unrespon-
sive, which contradicts the philosophy of DNAlive: the web must feel real-time
interactive and fast to the user.

4.3 The physical properties scripts


For a gene to express (e.g. Blue eyes), a protein (Transcription factor) attaches
to the DNA sequence associated with that gene (CATGACTGACGTACGTTAGC), and
1 In computational biology, a computing time of less than one minute is very fast.

27
Two types of inputs. The user can switch between them

Genomic DNA Calculate the physical


coordinates sequence 2D Plot
properties

Alternative: Work with a


DNA-Protein structure

UCSC Get Sequence


database Search for Proteins
proteins database

Genomes BLAT
database
Proteins

Draw 3D
Bind proteins
Structure

Calculate the physical


properties

DNA 3D DNA-Protein
Structure complex

Paint the
properties in the Animate DNA
3D structure

3D Plot 4D Plot

Figure 4.1: Architecture of the system. The user can input either genomic
coordinates or a DNA sequence, while being able to change between them at
any moment. The outputs (2D, 3D, 4D plot) are shown in blue, and all the
data structures are painted in red. Databases are painted in yellow, and all the
other processes are depicted in white.

28
then produces another protein. These proteins do not actually bind to a spe-
cific DNA sequence, but to a sequence with specific physical properties or,
computationally speaking, its attributes.
In Chapter 3 we presented a list of different works on DNA descriptors. For
this phase we chose 29 of them, described in the tables in appendix B.
Some examples of physical properties are:

• The melting temperature. Some parts of the DNA are more easily melted
than others; just like the edge of a paper burns more easily than the middle
part.
• Twist. The DNA can be physically twisted more easily in some specific
areas.
• Nucleosome preference. The probability that this DNA segment will bind
to a big protein named Nucleosome (see section 2.5)

AT TA GC TT GG CA AC
Melting temperature 0.023 0.102 0.073 0.009 0.069 0.104 0.044
Twist 0.033 0.017 0.025 0.026 0.026 0.016 0.036
Nucleosome preference 0.012 0.038 0.081 0.011 0.012 0.021 0.054

Table 4.1: Sample values for different physical properties

So, how can these properties be calculated? There are various methods, but
the main idea is to compute the value for each couple of nucleotides, and then
average them if needed. In table 4.1 we can see how the DNA “alphabetical”
sequence can be translated to plain numbers.
Let us imagine that we have the following DNA sequence:
CGTTAATATCGGCGTAGCTAGTGTTTTTCCGATATATCAGTCCGGGCCGCG
That sequence can be interpreted as follows:
XXXPPPPPPPPPPXXXXGGGGGGGGGGGGGGGGGGGGGGGGGGGGGGXXXX
being P a promoter, G a gene, and X nucleotides with no useful information.
We can transform this sequence to plain numeric values (again, see table
4.1) and, if needed, plot a histogram of the resulting values, as shown in figure
4.2

Figure 4.2: Plot of the “A-Philicity” property for a given sequence

Most of the descriptors use the Nearest Neighbor (NN) model. This approach
assumes that the stability of a given base pair depends on the identity and
orientation of neighboring base pairs.
The energy values are retrieved experimentally, in vitro. Scientists calculate
the energy values for 500 DNA sequences [BD98] and, then, using a linear
equation system, the values for each couple of base pairs is determined.

29
Once the energy values for each couple of base pairs and each property is
calculated, the algorithm needs to read the DNA sequence, aggregate the values
every couple of base pairs, and apply some correction factors. Usually, a sliding
window of 25 base pairs is used, so that every 25 bases the value is averaged and
the signal noise is reduced. These algorithms are, in fact, physical properties
predictors, as they can only estimate the actual value of the descriptors with
a mathematical model which is not perfect. However, this approach is widely
used and drops results which are very similar to the experimental ones.
To implement the scripts we used the Perl programming language. The main
reason is that there are already plenty of libraries for bioinformatic use. It also
interacts easily with web pages, and each script could be transformed into a web
page if needed.

4.4 Web page


The web page should be the main interface for DNAlive, as many users know
how to use a web browser. It needs to be easy to use, provide many help links
and guide the user through the process.
The technology choice was clear: HTML+CSS with Javascript for the page,
and PHP for the server-side scripts. Our server uses Apache with PHP5, which
is a very stable environment and allows object oriented programming. Our
bad experiences with Java web servers (see section 4.5.2) made us discard the
possibility of using JSP pages.
A PHP page was implemented for every major function, using Javascript
and hidden frames to exchange information between the pages. The main page
would call the helper pages when needed, retrieve the information, and call the
next subprocess. For the scripts side, we used temporal folders on the server for
exchanging information. A random token is generated every time that the user
reloads the page, allowing the server to centralize all the data on a single folder
and also providing a cache mechanism.
For the design details, the sample CSS from the MMB group pages
(mmb.pcb.ub.es) was used, with slight changes to improve readability. The
Scriptaculous2 Javascript library helped on the page transitions and provided a
beautiful and professional touch to the user interface.
Based on similar bioinformatics pages, we initially designed a very long form
where the user could input a lot of information. Unfortunately, the interface
was crowded with options and it was changed to a step-by-step guided process
(more in sec. 4.5.2). We then took advantage of the clean, new code and
designed a logging system and a development-production dual system based on
best-practices advices from [Hen06].
The logging system keeps track of any unexpected outputs or errors and —if
the error is fatal— redirects the user to an informative page announcing that
we are aware of the problem, and then we get a notification. This way, we could
track many hidden bugs and solve strange problems.
The development-production system keeps always an immutable production
version, where the beta testers discover new errors on a frozen environment,
while allowing the developers to work on new features and fix existing bugs. The
testers can then be notified when there is a new release version and check again
2 http://script.aculo.us/

30
if the bugs where fixed and provide feedback on the interface changes. This
system also keeps backups of previous versions and deploys the development
page to production instantly, so no downtime is noticed.

4.5 Web services


The Instituto Nacional de Bioinformática is one of the main providers for
BioMOBY web services3 . The services are scattered through the many
INB nodes, but the repository is hosted at the INB server.
BioMOBY [WL02] is an open source research project which aims to gener-
ate an architecture for the discovery and distribution of biological data through
web services; data and services are decentralized, but the availability of these
resources, and the instructions for interacting with them, are registered in a
central location called MOBY Central. BioMOBY adds to the web services
paradigm, as exemplified by Universal Data Discovery and Integration (UDDI),
by having an object-driven registry query system with object and service on-
tologies. This allows users to traverse expansive and disparate data sets where
each possible next step is presented based on the data object currently in-hand.
Moreover, a path from the current data object to a desired final data object
could be automatically discovered using the registry. Native BioMOBY objects
are lightweight XML, and make up both the query and the response of a Simple
Object Access Protocol (SOAP) transaction.
After analyzing the features of the system, we decided to implement all the
original algorithms from the web page into BioMOBY web services. We also
provide other services so that a complete workflow can be launched. Workflows
are nothing more than web services whose inputs and outputs are connected and
can generate many different outputs from just one input.
BioMOBY web services require very specific software versions and libraries.
They need to be implemented in Java 1.5 and jBoss 4.0.3sp1, as the most recent
versions of these software have changed the XML-RPC APIs and BioMOBY
does not support them yet. Nevertheless, it is an open platform, and there
is already some people at the INB working on a new API. The current one,
however, can be obtained via SVN, and then compiled and deployed to the
jBoss application server.
To develop new web services we used Taverna [KSW06], a tool that auto-
mates the process. First of all, the new service needs to be registered in the
yellow pages (in our case, the INB server), including all the input and output
data structures, the description, and so on. After that, Taverna generates the
skeleton for the web service, which needs to be imported into the development
IDE of choice —we used Netbeans—; then, it can be programmed as a standard
web service. The skeleton and the endpoint call the API when needed but, in
practice, we needed to perform some tweaking of the generated skeletons.
Two sets of web services were implemented: a workflow to calculate all
physical properties for a given DNA sequence (fig. 4.3), and another workflow
to generate annotations on a sequence (fig. 4.4).
3 http://www.inab.org/MOWServ/

31
Figure 4.3: Physical properties web services

Figure 4.4: Structural web services

32
4.5.1 Physical properties services
There is no need to dig deep into the implementation of these services, because
they are only a wrapper for the scripts. They receive the input data in the
standard BioMOBY format, call the scripts, reformat the output into BioMOBY
data, and return it to the user.
The services are grouped by descriptor types: DNA Flexibility, PARMBSC0
helical force constants, DNA Disruption energy, DNA non-linear dynamics,
DNA 3DNA structure, DNA stability, Unusual DNA conformation, ProStar
and Nucleosome potential4 .
The user can build a simple workflow, like the one in fig. 4.3, and launch
batch calculations for different DNA sequences, automating the process.

4.5.2 Structural services


The structural services allow the user to perform intermediate calculations or
just use our algorithms without an internet browser. The equivalences between
the architecture of the system (fig. 4.1) and the structural services workflow
(fig. 4.4) are:

• The service “showPDBFromFasta” converts a DNA sequence into a 3D


structure, in the PDB format5 . This implements the use case “Draw 3D
structure”

• The DNA-Protein use case (architecture: “Search for proteins”, “Bind


proteins”) is carried out by “runSimpleTfbs”, which searches for proteins;
“getPDBFromTransfac”, which retrieves the proteins from the database;
and “runBindProteins”, which finally binds the proteins to the DNA.

• There is one new use case: “show3DFeature”, whose inputs are a DNA
3D structure and information about the CpG islands (“runEmbossCp-
GReport GFF FromFasta”). This service outputs the DNA and a script
to draw the CpG information in the VMD software. The user then needs
to open this software, open the DNA molecule and paste the script code.
• Another service is provided: “runEmbossTFscan GFF FromFasta”, which
outputs information about the transcription factors on a DNA. It is a
wrapper for the program tfscan from EMBOSS, a bioinformatics soft-
ware package. The results are similar to those of “runSimpleTfbs”, and is
offered as an alternative to it.

Looking again at the system architecture, there are some process which have
no equivalent as a web service. First of all, we decided not to implement those
which are already provided as a web service by somebody else. In this case,
“BLAT” and “Get Sequence”.
Animating a DNA is also a web-only feature. We think that it is important
to offer services which can be useful to run as a batch, but watching a DNA
movie is an interactive process and it makes no sense to return a long file. With
4 In the web page, these two services appear as “Regulation parameters”, but we decided

to separate them to publicize ProStar, letting it appear as a single service.


5 PDB is the de-facto standard format for 3D DNA structures, and FASTA is a widely

extended format used to store sequences.

33
3D DNA it makes sense to return a PDB file, because they are widely used and
there is plenty of software to derive data from them.

4.6 Deployment and testing


Both interfaces were deployed on mmb2.pcb.ub.es, the MMB group’s applica-
tion server.
The web page stayed on a private beta phase for about five months, and
since very early stages many of the group members tested it frequently, thanks
to the dual development-production release cycle and the log system. The tests
were also very important for determining the final interface of the web, as the
first versions were hardly usable.
The implementation cycle for a web page is very interesting, as the testing
part accounts for most of the development time. It is composed of many mod-
ules, written in different languages, and bugs are difficult to track. There is also
the security problem, because a faulty input can open a breach in the server,
and all the parameters must be checked at least once in every layer (Javascript,
PHP, Apache).
A problem arose from one of the external software we use: Jmol, the 3D
molecular viewer, has very severe memory limitations. We investigated further
and discovered that Java Virtual Machines run very limited on memory. They
have a heap size of 64 megabytes for security reasons, because Java applets run
in client mode and could hang the system if the software is badly programmed.
That limits the size of the molecule that we can display and, furthermore, all
operations on the molecule cost a lot of CPU time, so we contacted the Jmol
team and found some workarounds to this issue. Still, at the time of writing
this document, the input for the modules which use Jmol is constrained, and
there is nothing that we can do.
Still another testing needs to be done, that is, browser compatibility. Our
group works with both Linux/Firefox and Mac/Safari machines, but Windows/IE
is still widely extended among researchers —and, more important, research
group leaders—. After much tweaking, we support IE6 or higher, Firefox2 or
higher, and Safari2 or higher, on all the platforms that run these browsers.
Opera 9+ also works, but it is not officially supported.
The web services were not tested exhaustively, as they do not really allow
the user much flexibility. If the input is valid, then the output will be correct.
Else, no output will be generated. Unfortunately, the 4.0.3sp1 version of jBoss
is not maintained anymore and has many bugs, so the main problems came from
the application server. Once we fixed most of them, the web services stay stable
and work fine.
To date, both the page and the web services have been working without an
issue.

4.7 Concluding remarks


The development of a large web application like DNAlive requires not only web
developing skills but also good planning and following an engineering process.

34
The web page needs to exchange data with external sites, thus using their avail-
able APIs.
Layered software design is essential in this case, since the data access and the
system logic need to be independent from the interface. Changing the algorithm
to compute a physical property, for example, cannot require modifications in the
interface code.
Finally, to develop research tools which can be computationally expensive,
one must also analyze the available hardware. Some of the scripts require expo-
nential algorithms, so all the batches need to be run through a queuing system,
to ensure hardware stability.

35
Chapter 5

Applying AI techniques to
ProStar

While Phase 1 was mainly devoted to the development of the DNAlive envi-
ronment, in this chapter we will present the new methodology used to improve
ProStar, whose original version was presented in section 3.5.
In order to follow the EGASP rules [BBB+ 06] and to be able to directly
compare the results with those of ProStar1, we got the same list of positive
promoter sequences and the set of non-promoter sequences as in the previous
version.
Positive sequences are the collection of 885 genes, annotated by the Havana
group in the ENCODE region (1% of the human genome) belonging to coding
genes. This information is stored on an index file havana.gene.
havana.gene contains the name of the gene, the name of the ENCODE
region where it belongs, the strand, the Transcription Start Site (TSS) and the
end of the transcribed region. A sample follows:
Havana1 ENm001 + 354319 390710
Considering the TSS as position 0, we selected the 2000 bases around it
[−1000.. + 1000] on the ENCODE database, resulting on 885 sequences, 2000
base pairs each. Further data calculations required us to drop 11 elements,
remaining 874 of them.
For the negative set, we got sequences composed of 2000 non-overlapping
bases (to genes and to themselves) lying in the same strand of transcribed
regions of the Havana genes. Using the example above, where 354319 is the
TSS, we retrieved the bases 353319 to 355318 for the positive set, and then, in
groups of 2000 bases, the rest of the data from 355319 to 390710, is added to
the negative set. Later on, we refined the negative set, deleting the elements
which were too close to a gene, as some of their properties were too similar to
those of the genes.
The strand information is also important. As the ENCODE database only
contains the positive strand, if the gene is located in the negative strand, the
resulting sequence needs to be complementary reversed. This means that AAGT
is reversed to TGAA and then complemented to ACTT.

36
5.1 Proposed solution
We are presented a classification problem which cannot be described by any
existent model. So, a new model needs to be discovered, and we propose to
learn it by using a neural network.
There are 29 DNA descriptors which we believe can be used to build this
model. However, we cannot use all of them to train the network, so we will
apply some filters beforehand, to the data set and also to the descriptors.
First, we will split the promoter database in four groups to gain specificity.
Then, an analysis of the DNA physical properties will be performed to reduce
the input noise on the network (fig. 5.2). After these techniques have been
applied, we will train a network for each group and study its performance as a
promoter classifier (fig. 5.3).

5.2 Promoter division in four groups


Promoters were introduced in section 2.3, along with their two key features:
TATA boxes and CpG islands. It has been widely proved [LGLP92] [SK03]
[MVDC+ 08] that these key elements, if present, affect the overall behavior of
the promoter.
Biologically, it is appropriate to distinguish those elements a priori, because
grouping promoters based on the presence of TATA/CpG can lead to more
accurate predictions.
A TATA-box related promoter is defined [Buc90] when, for a given 2000-
nts DNA sequence, with the annotated TSS in position +1000, there is at least
a 20-nts subsequence in the 9900-1000 range with a TATA-Score greater than a
cutoff value.
This TATA-Score is defined as the sum of the corresponding values for the
TATA weighted matrix, the most important part of which is depicted below. To
interpret it, read the DNA sequence in windows of n bases, being n the length
of the TATA matrix. Then, match each base to the corresponding value, and
sum all of them. If that sum is greater than the cutoff value, usually fixed at
-8.16, a TATA box is found.
As an example, the sequence GTATATAAG will have a TATA-box (0−0−0−0−
0−0.52−0−0−0 = −0.52 > −8.16), while the sequence AACCGGTTA will not have
one (−1.02 − 3.05 − 5.22 − 3.49 − 3.77 − 4.73 − 3.65 − 0.37 − 0 = −25.3 < −8.16).
Notice that the TATA box is very flexible in its sequence, as the example
CTTTGAATG, which one might not label as a TATA box at first sight, has a TATA-
score greater than the cutoff (−0.28 − 0 − 2.28 − 0 − 3.77 − 0 − 0 − 0.37 − 0 =
−6.7 > −8.16)
For a given 2000-nts DNA sequence, with the annotated TSS placed in the
position +1000, it will be considered associated to a CpG island if it contains
a region of greater than 200 bp with a percentage of G and C greater than 50%
(eq. 5.1), and the observed/expected ratio of CpG is greater than 0.6 (eq. 5.2)
[GGF87]. CpG refers to a C nucleotide immediately followed by a G. The “p”
in “CpG” refers to the phosphate group linking the two bases.
#C + #G
(5.1)
window length

37
Pos. -3 -2 -1 0 1 2 3 4 5
A -1.02 -3.05 0 -4.61 0 0 0 0 -0.01
C -0.28 -2.06 -5.22 -3.49 -5.17 -4.63 -4.12 -3.74 -1.13
G 0 -2.74 -4.28 -4.61 -3.77 -4.73 -2.65 -1.5 0
T -1.68 0 -2.28 0 -2.34 -0.52 -3.65 -0.37 -1.4

Table 5.1: Partial weight matrix for the TATA box, the full version can be found
at [Buc90]. Values for the “TATAA” sequence are rendered in bold.

#CpG × window length


(5.2)
#C × #G

5.2.1 TATA/CpG clustering


Both the positive and the negative set of sequences are clustered in 4 groups
using the aforementioned methods. The plus and minus symbols indicate if the
TATA box or CpG islands are present.

Pos Neg Total %Total


TATA- CpG- 407 4960 5367 75.6%
TATA- CpG+ 377 288 665 9.4%
TATA+ CpG- 60 959 1019 14.3%
TATA+ CpG+ 30 23 53 0.7%
Total 874 6230 7104 100.0%

Table 5.2: The dataset, divided into the four promoter groups.

This way, from now on, every cluster will have each its own positive and
negative set and will be treated independently.
The EGASP rules establish that 13 of the 44 ENCODE regions are used for
training and the others for testing. Unfortunately, after clustering the promoters
in four groups, a further division between a training set and a test set would
leave the smallest group empty, and the rest with few elements. Therefore it
was decided to use the whole dataset for training, performing a 10-fold cross-
validation.

5.3 DNA descriptor analysis


There are 27 physical descriptors of DNA to be included in the new version of the
promoter prediction, 26 of them belong to the 29 descriptors in section 4.3. We
believed that CpG islands are highly correlated with most physical properties,
so an extra descriptor for CpG islands was added1 . On the other side, the
Triplex and Quadruplex algorithms are very experimental and we might get
unexpected results, so they were discarded. ProStar1 was removed because it
1 This is the same concept as in sec. 5.2.1. Now, instead of using the official threshold for

deciding whether a CpG island is present, we will use all the resulting values as a descriptor.

38
is just a composition of the 6 helical properties, which are going to be analyzed
again.
We then run the descriptors on all the dataset, positive and negative. The
computing time took about 48 hours in 10 multi-core machines of the MMB grid
system; an estimated time of 1000 CPU-hours. Even though the process could
have been run through DNAlive web services, the naked scripts were used, so
that the software could be run through batch queues, avoid network and web-
services protocol overhead, and speed up the calculation time in general.
The descriptor analysis consists of two steps, depicted in figure 5.1. The
first step is to compute the Pearson correlation matrix for all descriptors, in
each promoter group. The results will show if there are significant differences
between promoter groups.
Then, a Principal Components Analysis per group was performed. The role
of the PCA is to gather the most important properties that define promoters,
and reduce the weight of those which do not add important information.

5.3.1 DNA parameter correlation


For every promoter type, we tried to narrow the descriptors to avoid overtrain-
ing our system by using parameters that do not add more information to the
promoter definition, so we studied the correlation of the properties.
The data now consists of 7104 elements (genes and non-genes, see table
above), with 27 descriptors for each. The descriptors are standardized arrays of
2,000 values, one for each nucleotide of the sequence.
In order to handle this amount of data, we decided to keep just 40 ele-
ments for each array. The average for each 50 elements was calculated, non-
overlapping, resulting in 2000/50 = 40 elements per array. Now each element
of the dataset has 27 ∗ 40 = 1080 descriptors, resulting in a more manageable
input matrix I of 7.6M items, but it was needed to reduce it even further to
feed the neural network (tab. 5.3).

Element cpg 50 cpg 100 cpg 150 ... z-dna 2000


Havana1 -0.649241 0.028227 0.028227 ... 0.705696
Havana5 -1.147745 -0.179814 -0.386870 ... -1.55970
...

Table 5.3: Sample descriptors matrix I before applying the eigenvectors func-
tion.

The first task was to run a Pearson correlation between all descriptors:

• For every promoter, every parameter is computed in 50-nts non-overlapping


windows, having 40 lectures per parameter per sequence.
• For every couple of physical properties, the Pearson correlation coefficient
is computed, ending with a 27x27 symmetric matrix.

• Finally, the Pearson correlation is calculated for each combination of prop-


erties and each promoter group.

39
Positive set

Compute the
physical Has TATA or CpG
properties

Physical
TATA/CpG
properties for
group
each sequence

Correlation
PCA
analysis

Correlation data
between Eigenvectors
properties

For each TATA/CpG group

Figure 5.1: Descriptor analysis methodology.

40
In general, there should be correlation between some physical properties,
either direct or inverse correlation. For example, the stability, by definition, is
inversely correlated to the free energy (fig. 6.5).
The results (fig. 6.4), indeed, show that some of them reach correlation
values even higher than 90% (colored in red in the aforementioned figures).
We decided then to run a PCA with all descriptors. As the algorithm builds
a correlation matrix to extract the eigenvectors, it should ignore (i.e. assign low
weights) to highly correlated descriptors and positions. Some of the blatantly
correlated properties could have been removed by hand, but we decided to test
the PCA analysis power by running it with all the data.

5.3.2 Principal components analysis


Pearson correlations are a useful first approach for analyzing data, but the
resulting information (see Results, 6.2.2) was not enough to train a neural net-
work. A Principal Components Analysis can automatically extract the
most relevant data of a big matrix, reducing all the variables to a combination
of eigenvectors.
Even though the negative set is about 7 times the size of the positive set,
the groups will be balanced to run the PCA. The conservation of information
was set to at least 80%.
Unfortunately, the initial results of the analysis showed a bias towards the
sequence position, as the first eigenvectors of the PCA were defined by sequence
positions and not the properties. For example, the position 1450 of the sequence
was always present in the first eigenvector for all the properties, then position
800 in the second eigenvector, and so on. That indicates a lot of noise caused by
the negative set and, while it is understandable and provides useful information
on the relevance of the sequence position for both sets, it was not what we
wanted.
Instead, we ran another PCA only on the positive set, getting better results.
In that case, eigenvectors were defined by properties with different positions,
which is more relevant for our problem.
The resulting eigenvectors look like Ei = weight× < property, position >.
Thus, a formula can be derived to transform the input data matrix I =<
property, position, value >, into a vector R with as many elements as eigen-
vectors has that promoter group.
X
R= I × Ei (5.3)
i

That way, we can transform the 1080 input parameters to only a few dozens
(Results, table 6.1), while conserving 80% of the original information. With less
parameters, the neural network to be trained (see next section) will be faster
and it should not overfit.

5.4 Predictor training


Once the descriptor analysis was performed, the next step consisted in train-
ing a neural network to classify the input sequences into promoters and non-
promoters. A neural network will be built for each promoter group, because their

41
signal for each physical property is different. Furthermore, the PCA reduced
the number of inputs, weighting the most important descriptors and ignoring
the others.
It is important to note that, even though we apply the positive set eigen-
vectors function to the input data, the neural network is fed with both sets to
provide a better profiling of each class. They were balanced in the following
manner:

Pos Neg Total %Total


TATA- CpG- 407 407 814 49.3%
TATA- CpG+ 377 288 665 40.3%
TATA+ CpG- 60 60 120 7.2%
TATA+ CpG+ 30 23 53 3.2%
Total 874 778 1652 100.0%

Table 5.4: Elements of the positive and negative set for each promoter group.
The positive set always kept all elements, dropping instances from the negative
set if necessary.

The inputs for the neural networks are the four vectors R (eq. 5.3). The
number of elements in each vector is the number of eigenvectors for that group.
Namely, 72 for the tata-cpg-, 69 for the tata-cpg+, 28 for tata+cpg- and
16 for tata+cpg+. As can be noticed, the number of eigenvectors increases
consistently with the number of elements in the group. Fig. 5.2 contains a
diagram with the data transformations that are performed in order to build the
classifier.
Weka 3.4.13 was used for this task, run in commandline mode. The classi-
fier was set to MultilayerPerceptron, and the other parameters were left with
the default values except: autobuild True, decay False, nominalToBinaryFil-
ter False, normalizeAttributes False, reset False. The training data was set to
cross-validate in 10 folds.
An initial combination of parameters for each network was obtained by brute
force. Combinations were computed by varying the learning rate between 0.1
and 1, step 0.1; the momentum between 0.1 and 1, step 0.1; and the number
of hidden layers between 1 and 3. That methodology allows us to assert that
our network quality is not far away from the best possible network for the given
data. Networks with more than 3 hidden layers were discarded because of the
computational cost and the marginal (if any) improvement in accuracy they
presented.
As the quality measure we looked at the absolute number of correctly classi-
fied instances. The results for these networks were analyzed and the best three
networks for each group were tweaked by hand, observing the impact of each
variable in the resulting. Besides these variables, also the number of neurons in
each layer and the training time were tested.
The four neural networks compose ProStar2’s classifier, acting as the class
predictor for promoter sequences. The models for each network can be found in
the results section (6.2.4).

42
Positive set Negative set

Eigenvectors

Apply the
eigenvectors
function
TATA/CpG
group

Model to define a Model to define a


promoter non-promoter

Neural network

NN
Promoter
classifier
For each TATA/CpG group

Figure 5.2: Data models for the predictor training.

43
5.5 ProStar 2.0
The new version of ProStar scans genomes in 2000-nts windows, moving forward
500-nts at each step. For every window we assume that the TSS would be placed
in position +1000, the position where the TSS was located in the training set.
We first classify this target sequence as CpG+/CpG- and TATA+/TATA-.
Once the promoter type is assigned to our target, we compute the appropriate
physical descriptors with DNAlive scripts. Then the eigenvectors function is
applied, and the resulting array is classified with the specific Neural Network.
A graphical image of this process is depicted in fig. 5.3.
ProStar 2 only requires the user to input a DNA sequence, and it automat-
ically performs all the calculations. Afterwards, it outputs the predicted class
(promoter or non-promoter) and the confidence value for its prediction.
Our method is not available to predict the strand of the promoter, so the
software will add an extra prediction with a negative strand to every promoter,
doubling the amount of predictions.

5.6 Evaluation of the results


For evaluating the results, we looked at the results for the 10-fold cross-validation
of each neural network. Weka outputs many accuracy descriptors, and we used
the following:
• Correctly classified instances, which is the absolute accuracy, as it aggre-
gates all the correct predictions.
• True positive, True negative, False positive and False negative percentages,
as depicted in the confusion matrix.
• Relative absolute error . This measure is a ratio of the mean absolute
error of the learning algorithm over the mean absolute error obtained by
predicting the mean of the training data. In some cases, the percentage of
correctly classified instances can be high because the predictions for the
largest group are good, but smaller groups get worse predictions.
True Positives (TP) are promoters which were classified as such by the
network. False Positives (FP) are non-promoters which were classified as
promoters. False Negatives (FN) are promoters which were classified as non-
promoters, and True Negatives (TN) were non-promoters correctly classified
as such. The accuracy is defined in equation 5.4
TP + TN
(5.4)
#Instances
As ProStar 1 does not cluster promoters by TATA/CpG, we will compare
its global accuracy to each of our groups.

5.7 Concluding remarks


In this section we have introduced a methodology composed of many AI tech-
niques. First, in order to increase the specificity of the predictions, the pro-
moters were split in four groups. Then, to reduce input noise, we narrowed

44
Compute the
Input sequence physical
properties

Physical
properties

Physical
Has TATA or CpG properties for the
current window

Eigenvectors Apply the


TATA/CpG
for that eigenvectors
group
group function

Data model

NN model
for that Run the classifier
group

Promoter
prediction
Iterate on a 2000-bps window, step 500 bps

List of all
predictions

Figure 5.3: ProStar2 algorithm.

45
the number of descriptors with the aid of two statistical techniques. Finally, a
neural network was built for each group. Each network was fed with the same
number of instances for both classes, when possible.
As a result, the ProStar 2 software scans the input sequence in windows of
the same size as the training set, stepping every 500 bases. It then classifies each
window into the promoter or non-promoter class, and outputs each prediction
with the confidence value.
To evaluate the results, we will compare the performance of each of the four
networks to ProStar 1 (Results, table 6.4).

46
Chapter 6

Results

6.1 DNAlive
DNAlive has had a big impact in researchers, corroborated by the publication
of a paper and two oral communications in international conferences1 . The web
page is accessed every day, especially for crossing ProStar’s predictions with
gene information at the UCSC Genome Browser.
DNAlive was developed to give a complete description of the physical prop-
erties of genomic DNA in a simple way, thus providing data that can be easily
understood by non-structural experts. Among others, DNAlive allows the user
to:

• Determine potential correlations between genome annotations (such as


transcription start sites, exons, splicing sites, . . . ) and a battery of 29
physical descriptors of DNA,
• Find out the most stable 3D structure of long genome fragments using
sequence-dependent average helical parameters, and, when available, ex-
perimental structural data on DNA-protein complexes,
• Perform a dynamic analysis of chromatin fiber exploring the range of de-
formability sampled during trajectory and the possibility of the formation
of transient proteinprotein complexes, and
• Display structural parameters of DNA in the context of associated func-
tional features obtained form several public databases.

Figure 6.1 depicts a screen shot of a DNAlive session in progress, and ap-
pendix A contains the published version of the manuscript.

6.2 ProStar 2
Many steps were involved in the new predictor methodology for ProStar 2. The
next sections present the individual results for each of the phases.
1 DNAlive: A tool for a realistic representation of the DNA at a genomic scale. VII
Jornadas de Bioinformtica, Valencia, February 2007. Representing the genomic DNA. Grand
challenges in computational biology, Barcelona, June 2008

47
Figure 6.1: Screen shot of a DNAlive session. The user has selected a 3D view
of a DNA sequence, and the Curvature descriptor is painted in a blue-to-red
scale.

48
Figure 6.2: Plot for the CpG property on the positive (top) and negative (bot-
tom) sets. The images on the left are the average for all promoter groups, while
the images on the right have each group separately. A clear signal can be ap-
preciated for the promoters, while the negative set only contains noise. The
negative tata+cpg+ (pink) group has higher variations because it contains less
elements.

6.2.1 DNA descriptors signal analysis


Five plots were generated for each descriptor, one per promoter group and
another for the aggregate of all the groups. Then, this was repeated for the
negative set, only to check that there is only noise when the promoter is not
present.
Figure 6.2 illustrates the results for the CpG property, which are very in-
teresting. The aggregated signal shows a clear peak at the TSS position. Sur-
prisingly, when looking at the split graphs, it is the non-cpg ones which contain
this peak (the blue one and the red one).
A CpG island (lines CPG+) has a very specific definition and requirements
(sec. 5.2.1). The results imply that, even if there are no CpG islands present,
the CpG% content alone is a good indicator for the presence of a TSS (position
20 in the x-axis of the graph). Furthermore, one can infer that the presence of
a CpG island in a 2000-bps sequence might reveal a promoter, even if the large
CpG content is not exactly positioned near the TSS.
Like CpG, many physical properties have clear signals on the TSS. Another
example is the curvature (fig. 6.3), which is displayed for illustration purposes.
Almost every physical property has a clear peak at the TSS, but some properties
work better for some promoter groups.
These results are very promising, as we did not expect to get such clear

49
Figure 6.3: Plot for the Curvature property on the positive (top) and negative
(bottom) sets. The images on the left are the average for all promoter groups,
while the images on the right have each group separately. Again, a clear signal
can be appreciated for the promoters, but not in the negative set.

50
TSS signals with that many properties. With this data, we can conclude that
the DNA descriptors can carry promoter signals, and that this signal behaves
differetly for each promoter group.
After looking at the graphs, we performed the numerical analysis of the
signals.

6.2.2 Analysis of the physicochemical properties


The correlation analysis resulted in four tables, one per promoter group, dis-
playing the Pearson correlation matrix between properties. In these tables,
correlations higher than 90% appear in red, between 20% and 90% in yellow,
and between 0% and 20% in white. For simplicity, only one of them is included
in this document, arbitrarily selected, the one for the positive set tata+cpg-
group, in the figure 6.4.
Also, the tables for the negative set are not shown. That is because the
correlation values were very similar, with a variation range of 10% at most,
thus concluding that physical properties are independent of the presence of a
promoter. Furthermore, the tables for each promoter group are very similar,
even though there are small differences.
For example, crossing CpG islands (descriptor cpg) with the six helical pa-
rameters (krise ...ktwist), shows that some of the parameters are more cor-
related with the groups which have CpG islands (krise), and some with the
groups without them (ktilt).
Furthermore, at least half of the properties are highly correlated with the
rest: basestack, bendstiff, denaturation, dupldisrfreeen, duplstabfree-
en, meltingtm, nuclpos, propwtist, protdeform, stability, stacking, z-dna.
This means that their values are probably a combination of some of the other
properties, or just that the biological interpretation for all of them is very sim-
ilar.
Figure 6.5 contains an image with some of the correlations, and besides
the graph in the foreground, one can see many highly correlated plots in the
background.

6.2.3 Principal components analysis


Once we built the descriptors matrix for all the input sequences, the PCA
outputted a number of eigenvectors for each promoter group (tab. 6.1). Each
of them was configured to keep 80% of the initial information. As expected, the
larger the number of input elements, the higher the number of eigenvectors.

Group Elements Eigenvectors


tata+cpg+ 30 17
tata+cpg- 60 29
tata-cpg- 377 73
tata-cpg+ 407 70

Table 6.1: Resulting eigenvectors after the PCA

Table 6.2 contains part of the first three eigenvectors for the tata-cpg+
group. In general, the first three eigenvectors carry about 20% of the original

51
Figure 6.4: Pearson correlation matrix for the TATA+ CpG- group.
52
Figure 6.5: Weka plotting the correlation between Stability and Duplex Stability
Free Energy.

information, and each one is a combination of weighted DNA properties/location


(sec. 5.3.2). The first 10 items usually carry between 6% and 7% of the total
weight, and they share the sequence position.

# Weight Elements
1 10.478% 6.4% ∗ denaturation1000 − 6.4% ∗ stacking1000 +
+6.3% ∗ meltingtm1000 − 6.3% ∗ duplstabf reeen1000 + . . .
2 8.204% −6.4% ∗ denaturation1850 − 6.4% ∗ meltingtm1850 +
+6.4% ∗ stacking1850 − 6.4% ∗ stability1850 + . . .
3 4.026% −6.7% ∗ meltingtm800 − 6.7% ∗ stability800 +
+6.7% ∗ zdna800 + 6.6% ∗ stacking800

Table 6.2: The first three eigenvectors extracted from the tata-cpg+’s PCA

6.2.4 Neural network


Table 6.3 illustrates the variables that were used to build the neural network
models. The methodology is detailed in section 5.4.
Results are evaluated as defined in sec. 5.6. Table 6.4 presents the accuracy
results for each of ProStar2’s promoter groups, versus ProStar1’s global accu-
racy, as the latest does not cluster the promoters. ProStar1’s accuracy (69%)
has been improved only for the tata+cpg+, which predicts promoters with an

53
Group Neurons / HL Learning rate Momentum Training time
tata+cpg+ 17, 17 0.5 0.2 500
tata-cpg+ 70 0.2 0.1 500
tata-cpg- 73 0.2 0.4 500
tata+cpg- 29, 29, 29 0.3 0.4 500

Table 6.3: Neural network models for each promoter group. The groups are
sorted by their accuracy (see tab. 6.4). The second column “Neurons per
hidden layer” represents both the number of layers and the neurons in each one,
separated by a comma. Note: in all cases, the number of neurons corresponds
to the number of eigenvectors for that group.

accuracy of 79%. The other groups’ accuracy is near 60% with very high values
for the relative error.
The relative absolute error (column RA Error) is about 50% in the best
case, and near 80% in the rest. This means that, in the worst case, the prediction
power is only a bit better than a ZeroR prediction (i.e. the mode of all values).

Group # Instances TP FP FN TN Accuracy RA Error


tata+cpg+ 53 25 6 5 17 79.2453% 52.4784%
tata-cpg+ 665 259 116 118 172 64.8120% 74.6566%
tata-cpg- 814 254 153 164 243 61.0565% 80.1061%
tata+cpg- 120 36 24 26 34 58.3333% 85.8145%
ProStar 1 1652 671 311 201 467 68.8861% Unknown

Table 6.4: Accuracy results for ProStar 2 versus ProStar 1 . The first column
indicates ProStar2’s promoter group, and the aggregated for ProStar 1. Pro-
Star1’s results were gathered with a different methodology, which did not output
the RA Error.

6.2.5 Final software


ProStar 2 is available as a web page at http://mmb2.pcb.ub.es/proStar2. It
is quite simple and it did not require any engineering process like DNAlive.
Instead, to emphasize the one-step prediction process, we designed a very min-
imalist interface. A screen shot of a ProStar 2 session is depicted at 6.6.
The interface is easy to use and no additional software is required other than
a web browser. The web page launches ProStar 2 in the MMB group application
server, and outputs the prediction when it is done.
The predictions of ProStar2 are not yet included in DNAlive, as we want
to keep the currently published version of the page until it undergoes a major
upgrade. No web service is available, for the same reason.

54
Figure 6.6: Screen shot of a ProStar 2 session. The user has input a DNA
sequence and the software has predicted a promoter at the position 1500.

55
Chapter 7

Conclusions

In chapter 1.1 we presented the dual objectives for this thesis: creating a multi-
interface platform to compute DNA descriptors, and implementing a new version
for the ProStar promoter predictor, using AI techniques.
In the next two sections we will analyze the results for each project and
compare them in detail to the expected objectives.

7.1 DNAlive
Huge databases are of no use without graphical tools to visualize the data and
modular algorithms that can process jobs in batches.
Cross-referencing data is a very difficult work, sometimes even impossible,
and at the current data growth rate researchers are now realizing that this
might be the last opportunity to classify it, now that we still have the required
computational power.
DNAlive was meant to be a simple interface for DNA description algorithms,
but it has turned out to be a step forward into building tools whose purpose is to
help users manage the available information. The implementation of the physi-
cal properties was derived from published work, data sources from the internet
are used to cross-reference data, and if it is available, we use the annotated
data from the UCSC Genome Browser to plot our tracks. 3D molecules and
animations are displayed in a Jmol applet, and our web services call other web
services to retrieve data and launch calculations.
To sum up, even if DNAlive has little Artificial Intelligence in it, it turns to
be an interesting project for the degree in Computer Engineering, and surely
will have more impact than ProStar 2.

7.2 ProStar 2
ProStar 2 has undergone a lot of changes from ProStar 1. Analyzing the method-
ology, it can be considered a different software rather than an evolution of the
first one. The three-step process has forced us to make some decisions, which
undoubtably have influenced the final results.
Some conclusions can be drawn from the obtained results:

56
• Promoter division in four groups. TATA boxes and CpG islands are the
most popular features for defining promoters, and the resulting clusters’
physical properties behaved different. This might mean that these features
are good for promoter clustering, although maybe not the best ones.
• Sequence extraction. We arbitrarily decided to select the 2000 nucleotides
surrounding each TSS. Looking at the signal for many properties, it is
clear that selecting about 1000-nts sequences might have reduced noise,
as the 500 initial and final nucleotides seem to carry no signal.
• Descriptors signal analysis. Most of the descriptors have a very clear
peak at the TSS. However, not all promoter groups benefit from the same
descriptors. In the Results chapter we described how the tata+cpg- group
had a very distinctive signal for the CpG descriptor, but this signal was
not that clear for the Curvature descriptor.
• Parameter correlation. In order to be computationally useful, 2000-nts
sequences were reduced to 27 descriptors of 40 values. We do not believe
that there is much information loss, as nearby nucleotides have similar
physical behaviors. Despite the easily visible correlations, all the descrip-
tors were used to train the neural network.
• Principal components analysis. We let the PCA decide the weight for each
descriptor and position, hoping that it would be smart enough to select
the most relevant ones. Looking at the resulting eigenvectors by hand,
some positions far away from the TSS —that is, with no relevant signal—
weighted more than what was expected.
• Predictor training. A MultilayerPerceptron was used as a classifier, and
because it is very versatile, after analyzing the results, it seems that this
was not a bad choice. Nonetheless, they tend to overfit and they are
difficult to tune. Again, we performed a brute force initial approach, and
the final results have been manually tuned based on the best networks.
However, as might have been expected, the networks with less inputs
didn’t work better (see table 6.4).

Having descriptors which plot such clear signals, we should think why Pro-
Star2’s predictions are not that accurate. We wanted to automate the weight
calculus for each descriptor with the PCA, so we did not filter them manually,
exchanging automation for accuracy. Parameters should have been filtered by
hand, keeping only the descriptors which had a clear signal (see sec. 6.2.1).
Still, we believe that the methodology presented in this thesis can be fine tuned
in order to increase other groups’ accuracy.
For all that, we think that the area which can be more improved on is
the Principal Components Analysis. The PCA cannot be blindly trusted to
reduce noise. Instead of letting it extract 72 eigenvectors from 1,080 posi-
tion/descriptors, it would be better to reduce the descriptors to about 100 and
then get more specific eigenvectors. Another approach to reduce the number
of eigenvectors is to keep less than 80% of the original information, say, about
50%.
The initial objective for ProStar2, improving ProStar1 predictive power, has
been reached only for the small tata+cpg+ promoter group (table 6.4). This

57
also suggests that clustering the promoters beforehand improves accuracy, but
another clustering based on different properties could be more appropriate.

7.3 Future work


DNAlive can be updated when there are significant changes to the DNA descrip-
tors. Also, it will adopt new technologies that allow researchers to visualize more
data in a web browser, for example, a Flash-based molecular viewer instead of
a Java one.
Furthermore, a new version is planned to integrate cross-species information
from the Evolutionary Genomics1 group at the IMIM.
Regarding ProStar 2, we have presented how promoter detection is a very
complex problem that cannot be tackled with simple mathematical functions
on energy values or text stemming techniques.
Our next steps will follow the ideas presented above. First of all, we will
analyze the descriptors’ correlations and signals per group, repeating the PCA
with less data and creating new neural networks with less input parameters.
Then, if the results do not improve, we will try to skip the PCA step and
manually create a set of weighted inputs for the network.
We are reaching a point where there is too much data available, and re-
searchers do not know where it can fit. Brute force techniques are computation-
ally impossible, and blind searches do not work. Machine learning algorithms
drop good results, but they need strong noise filters before they can be fed with
data.

1 http://evolutionarygenomics.imim.es/

58
Bibliography

[ABLC98] J F Allemand, D Bensimon, R Lavery, and V Croquette. Stretched


and overwound DNA forms a pauling-like structure with exposed
bases. Proc Natl Acad Sci USA, 95(24):14152–7, Nov 1998.
[AD07] P Akan and P Deloukas. DNA sequence and structural properties
as predictors of human and mouse promoters. Gene, Jan 2007.
[ASB+ 08] T Abeel, Y Saeys, E Bonnet, P Rouze, and Y Van de Peer. Generic
eukaryotic core promoter prediction using structural features of
DNA. Genome Res, Jan 2008.
[ASRdP08] T Abeel, Y Saeys, P Rouze, and Y Van de Peer. ProSOM: Core
promoter prediction based on unsupervised clustering of DNA
physical profiles. Bioinformatics, page 9, May 2008.
[BBB+ 06] Vladimir B Bajic, Michael R Brent, Randall H Brown, Adam
Frankish, Jennifer Harrow, Uwe Ohler, Victor V Solovyev, and
Sin Lam Tan. Performance assessment of promoter predictions
on ENCODE regions in the EGASP experiment. Genome Biol, 7
Suppl 1:S3.1–13, Jan 2006.
[BD98] R Blake and S Delcourt. Thermal stability of DNA. Nucleic Acids
Research, 26(14):3323–3332, 1998.
[BFBM86] K Breslauer, R Frank, H Blöcker, and L Marky. Predicting dna
duplex stability from the base sequence. Proceedings of the Na-
tional Academy of Sciences of the . . . , Jan 1986.
[BMHT91] A Bolshoy, P McNamara, R Harrington, and E Trifonov. Curved
DNA without AA: experimental estimation of all 16 DNA wedge
angles. Proceedings of the National Academy of Sciences of the
. . . , Jan 1991.
[BS03] Vladimir B Bajic and Seng Hong Seah. Dragon gene start finder:
an advanced system for finding approximate locations of the start
of gene transcriptional units. Genome Res, 13(8):1923–9, Aug
2003.
[BSC+ 02] Vladimir B Bajic, Seng Hong Seah, Allen Chong, Guanglan
Zhang, Judice L Y Koh, and Vladimir Brusic. Dragon promoter
finder: recognition of vertebrate RNA polymerase II promoters.
Bioinformatics, 18(1):198–9, Jan 2002.

59
[BSC+ 03] Vladimir B Bajic, Seng Hong Seah, Allen Chong, S P T Krishnan,
Judice L Y Koh, and Vladimir Brusic. Computer model for recog-
nition of functional transcription start sites in RNA polymerase II
promoters of vertebrates. J Mol Graph Model, 21(5):323–32, Mar
2003.
[BSSP95] I Brukner, R Sanchez, D Suck, and S Pongor. Trinucleotide mod-
els for DNA bending propensity: comparison of models based on
DNaseI digestion and . . . . J Biomol Struct Dyn, Jan 1995.

[BTC+ 06] Vladimir B Bajic, Sin Lam Tan, Alan Christoffels, Christian
Schönbach, Leonard Lipovich, Liang Yang, Oliver Hofmann,
Adele Kruger, Winston Hide, Chikatoshi Kai, Jun Kawai, David A
Hume, Piero Carninci, and Yoshihide Hayashizaki. Mice and men:
their promoter properties. PLoS Genet, 2(4):e54, Apr 2006.

[Buc90] P Bucher. Weight matrix descriptions of four eukaryotic RNA


polymerase II promoter elements derived from 502 unrelated pro-
moter sequences. J Mol Biol, 212(4):563–78, Apr 1990.
[Cea05] P Carninci and et al. The transcriptional landscape of the mam-
malian genome. Science, 309(5740):1559–63, Sep 2005.

[CLH+ 96] P Cluzel, A Lebrun, C Heller, R Lavery, J L Viovy, D Chate-


nay, and F Caron. DNA: an extensible molecule. Science,
271(5250):792–4, Feb 1996.
[Cri70] F Crick. Central dogma of molecular biology. Nature,
227(5258):561–3, Aug 1970.
[CW53] F Crick and J Watson. A structure for deoxyribose nucleic acid.
Nature, Jan 1953.
[Dav03] Ramana V Davuluri. Application of FirstEF to find promoters
and first exons in the human genome. Current protocols in bioin-
formatics / editoral board, Andreas D Baxevanis [et al], Chapter
4:Unit4.7, May 2003.
[DGZ01] R V Davuluri, I Grosse, and M Q Zhang. Computational identi-
fication of promoters and first exons in the human genome. Nat
Genet, 29(4):412–7, Dec 2001.

[DH02] Thomas A Down and Tim J P Hubbard. Computational detection


and location of transcription start sites in mammalian genomic
DNA. Genome Res, 12(3):458–61, Mar 2002.
[FSD+ 05] K Florquin, Y Saeys, S Degroeve, P Rouze, and Y Van de . . . .
Large-scale structural analysis of the core promoter in mammalian
and plant genomes. Nucleic Acids Research, Jan 2005.
[GB06] Samuel S Gross and Michael R Brent. Using multiple alignments
to improve gene prediction. J Comput Biol, 13(2):379–93, Mar
2006.

60
[GGF87] M Gardiner-Garden and M Frommer. CpG islands in vertebrate
genomes. J Mol Biol, Jan 1987.
[GPTO07] J Goñi, A Pérez, D Torrents, and M Orozco. Determining pro-
moter location based on DNA structure first-principles calcula-
tions. Genome Biology, Dec 2007.
[GT81] O Gotoh and Y Tagashira. Stabilities of nearest-neighbor dou-
blets in double-helical DNA determined by fitting calculated . . . .
Biopolymers, Jan 1981.
[GZW95] A Gorin, V Zhurkin, and K Wilma. B-DNA twisting correlates
with base-pair morphology. J Mol Biol, Jan 1995.
[HB05] Julian L Huppert and Shankar Balasubramanian. Prevalence of
quadruplexes in the human genome. Nucleic Acids Research,
33(9):2908–16, Jan 2005.
[HC96] M El Hassan and C Calladine. Propeller-twisting of base-pairs
and the conformational mobility of dinucleotide steps in DNA. J
Mol Biol, Jan 1996.
[Hen06] C Henderson. Building scalable web sites. page 330, Jan 2006.
[HS89] P A Harbury and K Struhl. Functional distinctions between yeast
TATA elements. Mol Cell Biol, 9(12):5298–304, Dec 1989.
[HZC90] P S Ho, G W Zhou, and L B Clark. Polarized electronic spectra
of Z-DNA single crystals. Biopolymers, 30(1-2):151–63, Jan 1990.
[IM94] V I Ivanov and L E Minchenkova. The A-form of DNA: in search
of the biological role. Mol Biol (Mosk), 28(6):1258–71, Jan 1994.
[KB05] Aditi Kanhere and Manju Bansal. A novel method for prokaryotic
promoter prediction based on DNA stability. BMC Bioinformat-
ics, 6:1, Jan 2005.
[Ken02] W Kent. . . . BLAT—The BLAST-like alignment tool. Genome
Res, Jan 2002.
[KKBB08] D Karolchik, R Kuhn, R Baertsch, and G Barber. The UCSC
genome browser database: 2008 update. Nucleic Acids Research,
Jan 2008.
[KSF+ 02] W James Kent, Charles W Sugnet, Terrence S Furey, Krishna M
Roskin, Tom H Pringle, Alan M Zahler, and David Haussler. The
human genome browser at UCSC. Genome Res, 12(6):996–1006,
Jun 2002.
[KSW06] Edward Kawas, Martin Senger, and Mark D Wilkinson. BioMoby
extensions to the taverna workflow management and enactment
software. BMC Bioinformatics, 7:523, Jan 2006.
[LGK05] Haibo Liu, Jianmin Gao, and Eric T Kool. Helix-forming proper-
ties of size-expanded DNA, an alternative four-base genetic form.
J Am Chem Soc, 127(5):1396–402, Feb 2005.

61
[LGLP92] F Larsen, G Gundersen, R Lopez, and H Prydz. CpG islands as
gene markers in the human genome. Genomics, 13(4):1095–107,
Aug 1992.

[LPKP01] V G Levitsky, O A Podkolodnaya, N A Kolchanov, and N L


Podkolodny. Nucleosome formation potential of eukaryotic DNA:
calculation and promoters analysis. Bioinformatics, 17(11):998–
1010, Nov 2001.

[MVDC+ 08] Vincent Miele, Cedric Vaillant, Yves D’aubenton-Carafa, Claude


Thermes, and Thierry Grange. DNA physical properties deter-
mine nucleosome occupancy from yeast to fly. Nucleic Acids Re-
search, page gkn262v1, May 2008.
[OcLNR02] Uwe Ohler, Guo chun Liao, Heinrich Niemann, and Gerald M Ru-
bin. Computational analysis of core promoters in the drosophila
genome. Genome Biology, 3(12):RESEARCH0087, Jan 2002.
[OGL+ 98] W Olson, A Gorin, X Lu, L Hock, and V Zhurkin. DNA sequence-
dependent deformability deduced from protein-DNA crystal com-
plexes. Proceedings of the National Academy of Sciences, Jan
1998.
[ONGR01] U Ohler, H Niemann, Liao Gc, and G M Rubin. Joint modeling
of DNA sequence and physical properties to improve eukaryotic
promoter recognition. Bioinformatics, 17 Suppl 1:S199–206, Jan
2001.

[ORBM78] R ORNSTEIN, R REIN, D BREEN, and R MACELROY. An


optimized potential function for the calculation of nucleic acid
interaction energies. I- Base . . . . Biopolymers, Jan 1978.
[OSHN00] U Ohler, G Stemmer, S Harbeck, and H Niemann. Stochastic
segment models of eukaryotic promoter regions. Pacific Sympo-
sium on Biocomputing Pacific Symposium on Biocomputing, pages
380–91, Jan 2000.
[PBCB98] A G Pedersen, P Baldi, Y Chauvin, and S Brunak. DNA structure
in human RNA polymerase ii promoters. J Mol Biol, 281(4):663–
73, Aug 1998.

[PMSS07] A Perez, I Marchan, D Svozil, and J Sponer. Refinement of the


AMBER force field for nucleic acids: Improving the description of
alpha/gamma . . . . Biophysical Journal, Jan 2007.
[RC96] R W Roberts and D M Crothers. Prediction of the stability of
DNA triplexes. Proc Natl Acad Sci USA, 93(9):4320–5, Apr 1996.

[San98] J SantaLucia. A unified view of polymer, dumbbell, and oligonu-


cleotide DNA nearest-neighbor thermodynamics. Proc Natl Acad
Sci USA, 95(4):1460–5, Feb 1998. El paper de meltingtm.
[SCH96] J Sun, T Carestier, and C Hélène. Oligonucleotide directed triple
helix formation. Curr Opin Struct Biol, Jan 1996.

62
[SCHH82] F Sanger, A Coulson, G Hong, and D Hill. Nucleotide sequence
of bacteriophage lambda DNA. J Mol Biol, Jan 1982.

[SDT86] S Satchwell, H Drew, and A Travers. Sequence periodicities in


chicken nucleosome core DNA. J Mol Biol, Jan 1986.
[SFL+ 95] M Sastry, R Fiala, R Lipman, M Tomasz, and D J Patel. Solution
structure of the monoalkylated mitomycin C-DNA complex. J
Mol Biol, 247(2):338–59, Mar 1995.

[SGLH97] J Sponer, H Gabb, J Leszczynski, and P Hobza. Base-base and


deoxyribose-base stacking interactions in B-DNA and Z-DNA: a
quantum-chemical study. Biophysical Journal, Jan 1997.
[SK95] A Sivolob and S Khrapunov. Translational positioning of nucle-
osomes on DNA: The role of sequence-dependent isotropic DNA
. . . . J Mol Biol, Jan 1995.
[SK03] Stephen T Smale and James T Kadonaga. The RNA polymerase
ii core promoter. Annual Review of Biochemistry, 72:449–79, Jan
2003.

[SKW00] M Scherf, A Klingenhoff, and T Werner. Highly specific local-


ization of promoter regions in large genomic sequences by pro-
moterinspector: a novel context analysis approach. J Mol Biol,
297(3):599–606, Mar 2000.
[SMHK01] Y Shen, K Musti, M Hiramoto, and H Kikuchi. Invariant asp-1122
and asp-1124 are essential residues for polymerization catalysis of
family D DNA polymerase from pyrococcus horikoshii. Journal of
Biological Chemistry, 276(29):27376–27383, Apr 2001.
[SNYH96] N Sugimoto, S Nakano, M Yoneyama, and K Honda. Improved
thermodynamic parameters and helix initiation factor to predict
stability of DNA duplexes. Nucleic Acids Research, 24(22):4501–
4505, 1996.
[SS03] V V Solovyev and I A Shahmuradov. PromH: Promoters identi-
fication using orthologous genomic sequences. Nucleic Acids Re-
search, 31(13):3540–5, Jul 2003.

[STS+ 01] Y Suzuki, T Tsunoda, J Sese, H Taira, J Mizushima-Sugano,


H Hata, T Ota, T Isogai, T Tanaka, Y Nakamura, A Suyama,
Y Sakaki, S Morishita, K Okubo, and S Sugano. Identification
and characterization of the potential promoter regions of 1031
kinds of human genes. Genome Res, 11(5):677–84, May 2001.

[VAM+ 01] J Venter, M Adams, E Myers, P Li, and R Mural. The sequence
of the human genome. Science, Jan 2001.
[vECLHP05] Titus S van Erp, Santiago Cuesta-Lopez, Johannes-Geert Hag-
mann, and Michel Peyrard. Can one predict DNA transcription
start sites by studying bubbles? Phys Rev Lett, 95(21):218104,
Nov 2005.

63
[VEH00] J M Vargason, B F Eichman, and P S Ho. The extended and
eccentric E-DNA structure induced by cytosine methylation or
bromination. Nat Struct Biol, 7(9):758–61, Sep 2000.

[WL02] Mark D Wilkinson and Matthew Links. BioMOBY: an open


source biological web services proposal. Brief Bioinformatics,
3(4):331–41, Dec 2002.

64
Index

Base, 7 False Negative, 44


Adenine, 7 False Positive, 44
Cytosine, 7 FirstEF, 20
Guanine, 7
Thymine, 7 Genetics, 9
Uracil, 8 Gene, 9
BioMOBY, 31 Gene expression, 10
Web services, 31 TSS, 9
BLAST, 19 Genome browser, 17

Central dogma of the molecular biol- Hidden Markov Models, 20


ogy, 6 Human genome, 3
Chromatin, 15
Chromosome, 15 Initiator element, 11
CpG island, 12, 37, 49
Mahalanobis distance, 21
DNA, 6, 7 Methylation, 12
3’ end, 7 Molecular dynamics, 19
3D Structure, 7 Monte Carlo method, 20
5’ end, 7
NSCAN, 20
A-DNA, 12
B-DNA, 12 Pearson correlation, 39
Backbone, 7 Pre-initiation complex, 10
Chemical Structure, 7 Principal Components Analysis, 41
Double helix, 7 Prokaryote, 6
Nucleosome, 15 promH, 20
Nucleotide, 7 Promoter, 3, 10
Physical properties, 29 Properties
Quadruplex, 13 Description, 73
Strand, 7 Plot, 29
Telomere, 14 Values, 29
Triplex, 13 ProStar 1, 21
Z-DNA, 12 Protein, 8
DNA Polymerase, 9 Polymerase II, 10
DNAlive, 26
Dragon Gene Start Finder, 20 Relative absolute error, 44, 54
Dragon Promoter Finder, 20 RNA, 8
mRNA, 10
ENCODE, 12
EP3, 24 TATA box, 11, 37
Eukaryote, 6 TATA score, 37

65
Taverna, 31
Transcription Factor, 10
True Negative, 44
True Positive, 44
TWINSCAN, 20

Workflow, 31

ZeroR, 54

66
Appendix A

DNAlive publication

In this appendix appears the publication for DNAlive and three pages of the
supplementary data where DNAlive is applied to find new genetic information.

67
Vol. 24 no. 15 2008, pages 1731–1732
BIOINFORMATICS APPLICATIONS NOTE doi:10.1093/bioinformatics/btn259

Structural bioinformatics

DNAlive: a tool for the physical analysis of DNA at the


genomic scale
J. Ramon Goñi1,2 , Carlos Fenollosa1,2,3 , Alberto Pérez1,2,3,4 , David Torrents1,2,5 and
Modesto Orozco1,2,3,4,∗
1 Joint
IRB-BSC Program on Computational Biology, Institute of Research in Biomedicine, Parc Científic de
Barcelona, Josep Samitier 1-5, Barcelona 08028, 2 Barcelona Supercomputing Center, Jordi Girona 31, Barcelona
08034, 3 National Institute of Bioinformatics, Parc Científic de Barcelona, Josep Samitier 1-5, 4 Departament de
Bioquímica, Facultat de Biología, Avgda Diagonal 647, Barcelona 08028 and 5 Institut Català per la Recerca i Estudis
Avançats (ICREA), Barcelona, Spain
Received on March 27, 2008; revised on May 16, 2008; accepted on June 4, 2008
Advance Access publication June 9, 2008
Associate Editor: Alfonso Valencia

ABSTRACT pairs away, or can affect dimerization of two DNA-binding proteins


Summary: DNAlive is a tool for the analysis and graphical display which might be separated by thousands of bases in sequence.
of structural and physical characteristics of genomic DNA. The web Different groups (Abeel et al., 2008; Goñi et al., 2007; Ohler et al.,
server implements a wide repertoire of metrics to derive physical 2001; Pedersen et al., 2000; Singhal et al., 2008) have demonstrated
information from DNA sequences with a powerful interface to derive that regulatory regions in DNA display unusual physical properties,
3D information on large sequences of both naked and protein-bound and in fact, two groups have recently proven independently (Abeel
DNAs. Furthermore, it implements a mesoscopic Metropolis code et al., 2008; Goñi et al., 2007) that eukaryotic promoters can
which allows the inexpensive study of the dynamic properties of be located with surprisingly good accuracy just analyzing simple
chromatin fibers. In addition, our server also surveys other protein physical descriptors of DNA, which confirms the existence of a
and genomic databases allowing the user to combine and explore hidden physical code that controls gene function. In summary,
the physical properties of selected DNA in the context of functional functional annotation needs to be complemented with physical data
features annotated on those regions. to understand the structure, dynamics and the general functionality
Availability: http://mmb.pcb.ub.es/DNAlive/ ; http://www.inab.org/ of genomic DNA.
Contact: modesto@mmb.pcb.ub.es DNAlive has been developed to give a complete description of the
Supplementary information: Supplementary data are available at physical properties of genomic DNA in a simple way, thus providing
Bioinformatics online. data that can be easily understood by non-structural experts.
Among others, DNAlive allows the user to (i) determine potential
correlations between genome annotations (such as transcription
1 INTRODUCTION start sites, exons, splicing sites, …) and a battery of 29 physical
Massive genomic projects have revealed the sequence of nearly descriptors of DNA (stability, helical descriptors, curvature, non-
50 eukaryotic genomes, including several mammals (among them, canonical B-DNA affinity, stiffness, …); (ii) find out the most
humans) and many more will become available in the coming stable 3D structure of long genome fragments (both naked DNA
years. So far, the annotation of these genomes has been nearly and DNA-protein complexes) using sequence-dependent average
restricted to the identification and the one-dimensional location helical parameters, and, when available, experimental structural data
of functional features (mostly genes and their regulatory regions), on DNA-protein complexes; (iii) perform a dynamic analysis of
without considering the structural parameters of their environment, chromatin fiber exploring the range of deformability sampled during
which have been proven to be crucial for the functionality of DNA. trajectory and the possibility of the formation of transient protein–
Determining the structural properties of DNA and the combination protein complexes and (iv) display structural parameters of DNA in
of functional features is necessary to interpret and understand the context of associated functional features obtained form several
the functionality of genomes in a more complex, and therefore public databases. The tool is available as a web page and also as
real, environment. The identification of these structural parameters different webservices, which can be incorporated in user workflows
allows scientists to consider different levels of accessibility of (Supplementary Material).
certain DNA regions to different proteins, such as transcription
factors, polymerases and DNA methylases. For example, specific
deformability or helical properties in a given region of DNA
2 IMPLEMENTATION
facilitate or impair the formation of nucleosomes hundreds of base 2.1 Entry data
The only mandatory input data for DNAlive is a DNA sequence

To whom correspondence should be addressed. in FASTA format or the genomic coordinates of a supported

© The Author 2008. Published by Oxford University Press. All rights reserved. For Permissions, please email: journals.permissions@oxfordjournals.org 1731

[12:03 18/7/03 Bioinformatics-btn259.tex] Page: 1731 1731–1732


J.R.Goñi et al.

vertebrate genome. The program retrieves parameters from their between conformation, functional annotations and physico-chemical
internal databases (Supplementary Table 1) to determine physical properties (Fig. 1).
profiles and to create a 3D structure of the naked DNA. The server also includes unique tools for a rapid representation
Given a DNA sequence, the program determines potentially of chromatin dynamics, which, in extensive analysis performed
bound transcription factor binding sites (TFBS) by scanning the in our laboratory on our database of more than 100 trajectories,
public TRANSFAC database (http://www.gene-regulation.com/) showed a surprisingly high accuracy of the essential deformation
linked to PDB (http://www.rcsb.org/) and Uniprot databases pattern of DNA. The method uses a mesoscopic Metropolis Monte
(http://www.ebi.uniprot.org/). The selection of the complex of Carlo algorithm, where the geometry of each base pair is defined
interest can be monitored externally by the user, who can force by three local rotations (roll, tilt and twist) and translations (slide,
the generation of specific complexes (for example, nucleosomes, shift and rise), and the conformational energy is estimated from the
protein-multicomplexes, etc.). deformation matrix using a harmonic model (Equation 1), where the
index ‘i’ stands for one of the M base pair steps and the index ‘j’
stands for the six unique helical parameters (ξ ) for each step. The
2.2 Server workflow
equilibrium values for one helical parameter in a given base pair step
Once a DNA sequence is entered (Fig. 1), the program computes type and (ξij0 ) and the associated deformation constant (Ki,j ) were
the profile for the 29 physical properties available for the fiber previously determined from molecular dynamics simulations (Pérez,
(Supplementary Table 1). All properties are represented in a 2D plot 2007). Once a movement in helical coordinates is accepted by the
using either the UCSC Genome Browser (http://genome.ucsc.edu) Metropolis test, the corresponding Cartesian structure of the fiber
in combination with annotated genes whenever genomic coordinates is generated using an adaptation of X3DNA (Lu and Olson, 2003)
for the genome are provided, or Gnuplot (Fig. 1 and Supplementary for VIDEO visualization using JMOL Java applets in the HTML
Fig. 1). page (Supplementary Fig. 3). Basic manipulation and analysis
To combine the visualization of DNA physical properties of the trajectories and structure (rotations, translations, distance
with public annotations of the genome, coordinates of the measurements,…) are allowed by the Jmol interface, which allows
input DNA sequence can be matched by running a search the determination of potential DNA-mediated protein-clusters.
in our local Blat server (Kent, 2002). Although the user is
able to annotate transcription factor PDB structures on specific 
M 
6  2
positions of the DNA input sequence, we have implemented E= Ki,j ξij −ξij0 (1)
an automatic method to perform this step using the TFBS Perl i=1 j=1
library (Lenhard and Wasserman, 2002). The reconstruction of the
average 3D structure of DNA is achieved using sequence-dependent ACKNOWLEDGEMENTS
base step parameters derived from accurate atomistic molecular
We thank the help of Agnes Noy, David Piedra, Henrique Proença
dynamics (Pérez, 2007) and making use of a local adaptation of
and Joaquín Panadero as β-testers of the server.
X3DNA (Lu and Olson, 2003) script (Fig. 1 and Supplementary
Fig. 2). When structural information on protein–DNA complexes Funding: This work has been supported by the Spanish Ministry
is available, modeled structures in the corresponding segment of Education and Science (BIO2006-01602 and BIO2006-
are substituted by the experimental geometries, and junctions 15036), the Spanish Ministry of Health (COMBIOMED network),
are refined if required. The visualization of 3D structures is the Fundación Marcelino Botín and the National Institute of
performed by integrating Jmol Java applets (http://www.jmol.org/) Bioinformatics (Structural Bioinformatics Node).
in the HTML page. All physical descriptors can be mapped into
the 3D structure to favor the detection of potential correlations Conflict of Interest: none declared.

REFERENCES
Abeel,T. et al. (2008) Generic eukaryotic core promoter prediction using structural
features of DNA. Genome Res., 18, 310–323.
Goñi,J.R. et al. (2007) Determining promoter location based on DNA structure first-
principles calculations. Genome Biol., 8, R263.
Kent,W.J. (2002) BLAT- the BLAST-like alignment tool. Genome Res., 12, 656–664.
Lenhard,B. and Wasserman,W.W. (2002) TFBS: computational framework for
transcription factor binding site analysis. Bioinformatics, 18, 1135–1136.
Lu,X.J. and Olson,W.K. (2003) 3DNA: a software package for the analysis, rebuilding
and visualization of three-dimensional nucleic acid structures. Nucleic Acids Res.,
31, 5108–5121.
Ohler,U. et al. (2001) Joint modeling of DNA sequence and physical properties
to improve eukaryotic promoter recognition. Bioinformatics, 17 (Suppl. 1),
S199–S206.
Pedersen,A.G. et al. (2000) A DNA structural atlas for Escherichia coli. J. Mol. Biol.,
299, 907–930.
Pérez,A. et al. (2007) Refinement of the AMBER force field for nucleic acids. Improving
the description of α/γ conformers. Biophys. J., 92, 3817–3829.
Singhal,P. et al. (2008) Prokaryotic gene finding based on physicochemical
characteristics of codons calculated from molecular dynamics simulations.
Fig. 1. DNAlive web server workflow diagram. Biophys. J. [EPub ahead of print; DOI:10.1529/biophysj.107.116392].

1732

[12:03 18/7/03 Bioinformatics-btn259.tex] Page: 1732 1731–1732






























Appendix B

Description of the physical


properties

The following tables contain the description for each of the 29 physical properties
and their references.

73
B.1 Unusual DNA conformation

Name Description Additional information Reference


Z-DNA Free energy values of Z- Derived from polarized electronic [HZC90]
DNA transition absorption spectra of single crys-
tals of Z-form duplexes
A-DNA Probability derived from Physochemical parameters de- [IM94]
energy cost of the B-DNA rived using the neighbor ap-
to A-DNA cost proach.
Triplex sta- Stability estimation of Parameters (enthalpies, Tm and [RC96]
bility DNA triple helices free energies) obtained assuming
neighbour approach in parallel
triplexes. Experimental values
derived from calorimetric stud-
ies of different triplex sequence at
different pHs.
Quadruplex Potential quadruplex se- Empirical method for predic- [HB05]
quence tion of G-DNA tetrads based on
the strand stoichiometry, num-
ber of stacked tetrads, muta-
tions/deletion and length and
composition of loops.

Table B.1: Unusual DNA conformation

74
B.2 DNA disruption energy

Name Description Additional information Reference


Base-pair Base-stacking energy Derived from simple force-field [ORBM78]
stacking energies using equilibrium ge-
ometries
Duplex dis- DNA disrupt energy Parameters (enthalpies and [BFBM86]
ruption free entropies) obtained assuming
energy neighbour approach. Exper-
imental values derived from
calorimetric studies of 19 DNA
oligomers and 9 DNA polymers.
Duplex sta- Thermodynamic free en- Corrected-near neighbour pa- [SNYH96]
bility free en- ergy rameters fit to experimental data
ergy (50) and cross-validated with
other 15 oligos. Values fitted to
melting curves.
Stacking en- Stacking energy from Derived from high quality quan- [SGLH97]
ergy quantum-chemical calcu- tum mechanical calculations (gas
lations phase) on equilibrium geometries
of the 10 unique DNA dimer
steps.

Table B.2: DNA disruption energy

75
B.3 DNA 3DNA structure

Name Description Additional information Reference


B-DNA twist Twist angle torsion based Geometrical parameters de- [GZW95]
on B-DNA crystals rived from analysis of crystal
databases (mostly slide and
propeller twist analysis). They
are obtained assuming neighbor
approach.
Curvature Curvature based on twist, Parameters derived from crystal [BMHT91]
wedge, direction calcula- structures of B-DNA and valid
tions from gel retardation within the neighbour limit.
experiments
Direction Direction of the deflection Parameters derived from crystal [BMHT91]
Wedge of the axial path of DNA structures of B-DNA and valid
within the neighbour limit.
Protein Twist angle torsion Geometrical parameters de- [OGL+ 98]
DNA twist based on Protein-DNA rived from analysis of crystal
complexes databases of DNA-protein com-
plexes within the neighbour
approach.
DNA Propeller-twist base pair Geometrical parameters de- [HC96]
propeller- measure from crystallo- rived from analysis of crystal
twist graphic data databases. Mobility is repre-
sented from deviations in the
propeller twist/slide profile.
Values are obtained assuming
neighbour approach.

Table B.3: DNA 3DNA structure

76
B.4 DNA flexibility

Name Description Additional information Reference


Bendability Deformability based on Parameters derved from DNase [BSSP95]
DnaseI cutting frequencies I digestion and nucleosome bind-
ing data and applied at the trin-
ucleotide level.
Bending Rigidity of the DNA helix Method to predict nucleosomal [SK95]
stiffness translational position in terms
of bending free energy computed
from the near neighbour model.
Nucleosome Nucleosome trinucleotide Parameters derived to predict [SDT86]
preference preference rotational preference of nucleo-
somes based on fitting to 177
sequences of chicken erythrocite
core particles.
Protein Deformability based on Deformation parameters de- [OGL+ 98]
deformation DNA-protein crystal rived from analysis of crystal
structures databases of DNA-protein com-
plexes within the neighbour
approach.

Table B.4: DNA Flexibility

B.5 DNA stability

Name Description Additional information Reference


Denaturation Denaturation equilibrium Experimental values derived [BD98]
Parameters (enthalpies from high-resolution melting
and entropies) obtained curves of synthetic DNAs
assuming neighbour inserted in pN/MCS plasmids.
approach.
Melting tem- Melting temperature Derived by fitting to melting pro- [GT81]
perature files of restriction fragment of
PHIx174 and fd phage DNAs.
Near neighbor model is used.
Stability DNA melting temperature Unified view of DNA melting. [San98]
derived from single set of Near neighbor parameters ob-
parameters tained by fitting accurate calori-
metric data.

Table B.5: DNA stability

77
B.6 DNA non-linear dynamics

Name Description Additional information Reference


Bubbles Non-linear predictor of Derived from Peyrard-Bishop- [vECLHP05]
bubble formation on DNA Dauxois dynamics simulations
using an ultrasimplified quasi-
harmonic potential fitted to re-
produce denaturalization curves
of short hereogeneous DNA seg-
ments.

Table B.6: DNA non-linear dynamics

B.7 PARMBSC0 Helical force constants

Name Description Additional information Reference


Rise Z-axis translational de- Derived from atomistic MD
formability simulations in water for a small
Roll Y-axis rotational deforma- set of duplexes cointaining all
[PMSS07]
bility unique dinucleotide steps.
Shift X-axis translational de- Helical force-constants are
formability derived by inversion of the
Slide Y-axis translational de- covariance matrix in helical
formability space and assuming harmonic
Tilt X-axis rotational deforma- oscillations. The neighbor
bility approach is used.
Twist Z-axis rotational deforma-
bility

Table B.7: PARMBSC0 Helical force constants.

78
B.8 Regulation parameters

Name Description Additional information Reference


ProStar Prediction of promoter re- Ab initio promoter prediction of [GPTO07]
gions using DNA-physical mammal genomes. The method
properties used helical force constants as
descriptors and is trained using
known annotated promoters in
humans.
Nucleosome Nucleosome preferential Derived by a Monte Carlo [LPKP01]
potential occupancy estimator code using nucleosome forma-
tion potentials obtained from
trained dinucleotide parameters.
Method was fit using a dataset of
141 nucleotide sequences which
were experimentally annotated.

Table B.8: Regulation parameters

79
This document was printed on September 3, 2008

You might also like