header logo image


Page 44«..1020..43444546..50..»

Archive for the ‘Genetic medicine’ Category

Master of Science in Genetic Counseling | Icahn School of …

Tuesday, November 29th, 2016

The Master of Science in Genetic Counseling Program at the Icahn School of Medicine at Mount Sinai (ISMMS) provides students with many opportunities for study, research, and practice in the challenging and exciting field of genetic counseling.

Our program, part of the ISMMS Graduate School of Biomedical Sciences, is sponsored by the Department of Genetics and Genomics Sciences, one of the largest departments of its kind in the world. Our faculty is renowned in the diagnosis, treatment and counseling of genetic disorders and is at the forefront of basic science research. Our large multidisciplinary center provides clinical and laboratory services to a wide range of culturally diverse patients and their families, along with our proven commitment to the community we serve.

The Master of Science in Genetic Counseling program is a 21-month, full-time course of study accredited by the Accreditation Council for Genetic Counseling (ACGC), which blends didactic coursework with a variety of clinical rotations.The Program has a 100% job placement rate for those seeking positions within the genetic counseling field. The first time Board certification pass rate for the classes of 2013, 2014 and 2015 is 95.5%.

Our students grow in skills and knowledge while enjoying access to the incomparable opportunities present in a prestigious academic medical center in the heart of New York City.

The Genetic Counseling Program was established in 1991 as a Certificate Program to train individuals with related advanced degrees to become genetic counselors. It became a Master of Science degree program in 1995 and is accredited by the Accreditation Council for Genetic Counseling (ACGC). As part of The Mount Sinai Health System.Our students have access to the laboratories, libraries, and educational resources of ISMMS as part of The Mount Sinai Health System.

The integration of academic and clinical disciplines within one of the country's preeminent health systems provides an ideal environment for our master's program, affording our students unparalleled opportunities for study, research, and practice in the challenging and exciting field of genetic counseling.

Randi Zinberg, MS, CGC Director, Master of Science Program in Genetic Counseling Email: randi.zinberg@mssm.edu

Sabrina Suckiel, MS, CGC Assistant Director, Master of Science Program in Genetic Counseling Email: sabrina.suckiel@mssm.edu

Julie McGlynn, MS, CGC Director, Clinical Training, Master of Science Program in Genetic Counseling Email: julie.mcglynn@mssm.edu

Genetics and Genomics Faculty

Who are genetic counselors?

Integrating the digital universe to better diagnose, treat and prevent disease

A professional network for the interests of genetic counselors

Promoting high standards in genetic counseling

Student Life and Resources

1st Year Genetic Counseling Program Experience at ISMMS

Live in one of the worlds most vibrant cities

Finance your graduate education

Course registrations, academic calendars, transcripts and more

Embracing the principles of diversity and inclusion

Learn about graduate student housing

Discover and explore research at ISMMS

Read the original:
Master of Science in Genetic Counseling | Icahn School of ...

Read More...

Genetics & Medicine – Site Guide – NCBI

Monday, November 7th, 2016

Bookshelf

A collection of biomedical books that can be searched directly or from linked data in other NCBI databases. The collection includes biomedical textbooks, other scientific titles, genetic resources such as GeneReviews, and NCBI help manuals.

A resource to provide a public, tracked record of reported relationships between human variation and observed health status with supporting evidence. Related information intheNIH Genetic Testing Registry (GTR),MedGen,Gene,OMIM,PubMedand other sources is accessible through hyperlinks on the records.

A registry and results database of publicly- and privately-supported clinical studies of human participants conducted around the world.

An archive and distribution center for the description and results of studies which investigate the interaction of genotype and phenotype. These studies include genome-wide association (GWAS), medical resequencing, molecular diagnostic assays, as well as association between genotype and non-clinical traits.

An open, publicly accessible platform where the HLA community can submit, edit, view, and exchange data related to the human major histocompatibility complex. It consists of an interactive Alignment Viewer for HLA and related genes, an MHC microsatellite database, a sequence interpretation site for Sequencing Based Typing (SBT), and a Primer/Probe database.

A searchable database of genes, focusing on genomes that have been completely sequenced and that have an active research community to contribute gene-specific data. Information includes nomenclature, chromosomal localization, gene products and their attributes (e.g., protein interactions), associated markers, phenotypes, interactions, and links to citations, sequences, variation details, maps, expression reports, homologs, protein domain content, and external databases.

A collection of expert-authored, peer-reviewed disease descriptions on the NCBI Bookshelf that apply genetic testing to the diagnosis, management, and genetic counseling of patients and families with specific inherited conditions.

Summaries of information for selected genetic disorders with discussions of the underlying mutation(s) and clinical features, as well as links to related databases and organizations.

A voluntary registry of genetic tests and laboratories, with detailed information about the tests such as what is measured and analytic and clinical validity. GTR also is a nexus for information about genetic conditions and provides context-specific links to a variety of resources, including practice guidelines, published literature, and genetic data/information. The initial scope of GTR includes single gene tests for Mendelian disorders, as well as arrays, panels and pharmacogenetic tests.

A database of known interactions of HIV-1 proteins with proteins from human hosts. It provides annotated bibliographies of published reports of protein interactions, with links to the corresponding PubMed records and sequence data.

A compilation of data from the NIAID Influenza Genome Sequencing Project and GenBank. It provides tools for flu sequence analysis, annotation and submission to GenBank. This resource also has links to other flu sequence resources, and publications and general information about flu viruses.

A portal to information about medical genetics. MedGen includes term lists from multiple sources and organizes them into concept groupings and hierarchies. Links are also provided to information related to those concepts in the NIH Genetic Testing Registry (GTR), ClinVar,Gene, OMIM, PubMed, and other sources.

A project involving the collection and analysis of bacterial pathogen genomic sequences originating from food, environmental and patient isolates. Currently, an automated pipeline clusters and identifies sequences supplied primarily by public health laboratories to assist in the investigation of foodborne disease outbreaks and discover potential sources of food contamination.

A database of human genes and genetic disorders. NCBI maintains current content and continues to support its searching and integration with other NCBI databases. However, OMIM now has a new home at omim.org, and users are directed to this site for full record displays.

A database of citations and abstracts for biomedical literature from MEDLINE and additional life science journals. Links are provided when full text versions of the articles are available via PubMed Central (described below) or other websites.

A digital archive of full-text biomedical and life sciences journal literature, including clinical medicine and public health.

A collection of clinical effectiveness reviews and other resources to help consumers and clinicians use and understand clinical research results. These are drawn from the NCBI Bookshelf and PubMed, including published systematic reviews from organizations such as the Agency for Health Care Research and Quality, The Cochrane Collaboration, and others (see complete listing). Links to full text articles are provided when available.

A collection of resources specifically designed to support the research of retroviruses, including a genotyping tool that uses the BLAST algorithm to identify the genotype of a query sequence; an alignment tool for global alignment of multiple sequences; an HIV-1 automatic sequence annotation tool; and annotated maps of numerous retroviruses viewable in GenBank, FASTA, and graphic formats, with links to associated sequence records.

A summary of data for the SARS coronavirus (CoV), including links to the most recent sequence data and publications, links to other SARS related resources, and a pre-computed alignment of genome sequences from various isolates.

An extension of the Influenza Virus Resource to other organisms, providing an interface to download sequence sets of selected viruses, analysis tools, including virus-specific BLAST pages, and genome annotation pipelines.

Read more:
Genetics & Medicine - Site Guide - NCBI

Read More...

DNA – Wikipedia

Friday, October 28th, 2016

Deoxyribonucleic acid (i;[1]DNA) is a molecule that carries the genetic instructions used in the growth, development, functioning and reproduction of all known living organisms and many viruses. DNA and RNA are nucleic acids; alongside proteins, lipids and complex carbohydrates (polysaccharides), they are one of the four major types of macromolecules that are essential for all known forms of life. Most DNA molecules consist of two biopolymer strands coiled around each other to form a double helix.

The two DNA strands are termed polynucleotides since they are composed of simpler monomer units called nucleotides.[2][3] Each nucleotide is composed of one of four nitrogen-containing nucleobaseseither cytosine (C), guanine (G), adenine (A), or thymine (T)and a sugar called deoxyribose and a phosphate group. The nucleotides are joined to one another in a chain by covalent bonds between the sugar of one nucleotide and the phosphate of the next, resulting in an alternating sugar-phosphate backbone. The nitrogenous bases of the two separate polynucleotide strands are bound together (according to base pairing rules (A with T, and C with G) with hydrogen bonds to make double-stranded DNA. The total amount of related DNA base pairs on Earth is estimated at 5.0 x 1037, and weighs 50 billion tonnes.[4] In comparison, the total mass of the biosphere has been estimated to be as much as 4 trillion tons of carbon (TtC).[5]

DNA stores biological information. The DNA backbone is resistant to cleavage, and both strands of the double-stranded structure store the same biological information. This information is replicated as and when the two strands separate. A large part of DNA (more than 98% for humans) is non-coding, meaning that these sections do not serve as patterns for protein sequences.

The two strands of DNA run in opposite directions to each other and are thus antiparallel. Attached to each sugar is one of four types of nucleobases (informally, bases). It is the sequence of these four nucleobases along the backbone that encodes biological information. RNA strands are created using DNA strands as a template in a process called transcription. Under the genetic code, these RNA strands are translated to specify the sequence of amino acids within proteins in a process called translation.

Within eukaryotic cells, DNA is organized into long structures called chromosomes. During cell division these chromosomes are duplicated in the process of DNA replication, providing each cell its own complete set of chromosomes. Eukaryotic organisms (animals, plants, fungi, and protists) store most of their DNA inside the cell nucleus and some of their DNA in organelles, such as mitochondria or chloroplasts.[6] In contrast, prokaryotes (bacteria and archaea) store their DNA only in the cytoplasm. Within the eukaryotic chromosomes, chromatin proteins such as histones compact and organize DNA. These compact structures guide the interactions between DNA and other proteins, helping control which parts of the DNA are transcribed.

DNA was first isolated by Friedrich Miescher in 1869. Its molecular structure was identified by James Watson and Francis Crick in 1953, whose model-building efforts were guided by X-ray diffraction data acquired by Rosalind Franklin. DNA is used by researchers as a molecular tool to explore physical laws and theories, such as the ergodic theorem and the theory of elasticity. The unique material properties of DNA have made it an attractive molecule for material scientists and engineers interested in micro- and nano-fabrication. Among notable advances in this field are DNA origami and DNA-based hybrid materials.[7]

DNA is a long polymer made from repeating units called nucleotides.[8][9] The structure of DNA is non-static,[10] all species comprises two helical chains each coiled round the same axis, and each with a pitch of 34ngstrms (3.4nanometres) and a radius of 10ngstrms (1.0nanometre).[11] According to another study, when measured in a particular solution, the DNA chain measured 22 to 26ngstrms wide (2.2 to 2.6nanometres), and one nucleotide unit measured 3.3 (0.33nm) long.[12] Although each individual repeating unit is very small, DNA polymers can be very large molecules containing millions of nucleotides. For instance, the DNA in the largest human chromosome, chromosome number 1, consists of approximately 220 million base pairs[13] and would be 85mm long if straightened.

In living organisms DNA does not usually exist as a single molecule, but instead as a pair of molecules that are held tightly together.[14][15] These two long strands entwine like vines, in the shape of a double helix. The nucleotide contains both a segment of the backbone of the molecule (which holds the chain together) and a nucleobase (which interacts with the other DNA strand in the helix). A nucleobase linked to a sugar is called a nucleoside and a base linked to a sugar and one or more phosphate groups is called a nucleotide. A polymer comprising multiple linked nucleotides (as in DNA) is called a polynucleotide.[16]

The backbone of the DNA strand is made from alternating phosphate and sugar residues.[17] The sugar in DNA is 2-deoxyribose, which is a pentose (five-carbon) sugar. The sugars are joined together by phosphate groups that form phosphodiester bonds between the third and fifth carbon atoms of adjacent sugar rings. These asymmetric bonds mean a strand of DNA has a direction. In a double helix, the direction of the nucleotides in one strand is opposite to their direction in the other strand: the strands are antiparallel. The asymmetric ends of DNA strands are said to have a directionality of five prime (5) and three prime (3), with the 5 end having a terminal phosphate group and the 3 end a terminal hydroxyl group. One major difference between DNA and RNA is the sugar, with the 2-deoxyribose in DNA being replaced by the alternative pentose sugar ribose in RNA.[15]

The DNA double helix is stabilized primarily by two forces: hydrogen bonds between nucleotides and base-stacking interactions among aromatic nucleobases.[19] In the aqueous environment of the cell, the conjugated bonds of nucleotide bases align perpendicular to the axis of the DNA molecule, minimizing their interaction with the solvation shell. The four bases found in DNA are adenine (A), cytosine (C), guanine (G) and thymine (T). These four bases are attached to the sugar-phosphate to form the complete nucleotide, as shown for adenosine monophosphate. Adenine pairs with thymine and guanine pairs with cytosine. It was represented by A-T base pairs and G-C base pairs.[20][21]

The nucleobases are classified into two types: the purines, A and G, being fused five- and six-membered heterocyclic compounds, and the pyrimidines, the six-membered rings C and T.[15] A fifth pyrimidine nucleobase, uracil (U), usually takes the place of thymine in RNA and differs from thymine by lacking a methyl group on its ring. In addition to RNA and DNA, many artificial nucleic acid analogues have been created to study the properties of nucleic acids, or for use in biotechnology.[22]

Uracil is not usually found in DNA, occurring only as a breakdown product of cytosine. However, in several bacteriophages, Bacillus subtilis bacteriophages PBS1 and PBS2 and Yersinia bacteriophage piR1-37, thymine has been replaced by uracil.[23] Another phage - Staphylococcal phage S6 - has been identified with a genome where thymine has been replaced by uracil.[24]

Base J (beta-d-glucopyranosyloxymethyluracil), a modified form of uracil, is also found in several organisms: the flagellates Diplonema and Euglena, and all the kinetoplastid genera.[25] Biosynthesis of J occurs in two steps: in the first step a specific thymidine in DNA is converted into hydroxymethyldeoxyuridine; in the second HOMedU is glycosylated to form J.[26] Proteins that bind specifically to this base have been identified.[27][28][29] These proteins appear to be distant relatives of the Tet1 oncogene that is involved in the pathogenesis of acute myeloid leukemia.[30] J appears to act as a termination signal for RNA polymerase II.[31][32]

Twin helical strands form the DNA backbone. Another double helix may be found tracing the spaces, or grooves, between the strands. These voids are adjacent to the base pairs and may provide a binding site. As the strands are not symmetrically located with respect to each other, the grooves are unequally sized. One groove, the major groove, is 22 wide and the other, the minor groove, is 12 wide.[33] The width of the major groove means that the edges of the bases are more accessible in the major groove than in the minor groove. As a result, proteins such as transcription factors that can bind to specific sequences in double-stranded DNA usually make contact with the sides of the bases exposed in the major groove.[34] This situation varies in unusual conformations of DNA within the cell (see below), but the major and minor grooves are always named to reflect the differences in size that would be seen if the DNA is twisted back into the ordinary B form.

In a DNA double helix, each type of nucleobase on one strand bonds with just one type of nucleobase on the other strand. This is called complementary base pairing. Here, purines form hydrogen bonds to pyrimidines, with adenine bonding only to thymine in two hydrogen bonds, and cytosine bonding only to guanine in three hydrogen bonds. This arrangement of two nucleotides binding together across the double helix is called a base pair. As hydrogen bonds are not covalent, they can be broken and rejoined relatively easily. The two strands of DNA in a double helix can thus be pulled apart like a zipper, either by a mechanical force or high temperature.[35] As a result of this base pair complementarity, all the information in the double-stranded sequence of a DNA helix is duplicated on each strand, which is vital in DNA replication. This reversible and specific interaction between complementary base pairs is critical for all the functions of DNA in living organisms.[9]

The two types of base pairs form different numbers of hydrogen bonds, AT forming two hydrogen bonds, and GC forming three hydrogen bonds (see figures, right). DNA with high GC-content is more stable than DNA with low GC-content.

As noted above, most DNA molecules are actually two polymer strands, bound together in a helical fashion by noncovalent bonds; this double stranded structure (dsDNA) is maintained largely by the intrastrand base stacking interactions, which are strongest for G,C stacks. The two strands can come apart a process known as melting to form two single-stranded DNA molecules (ssDNA) molecules. Melting occurs at high temperature, low salt and high pH (low pH also melts DNA, but since DNA is unstable due to acid depurination, low pH is rarely used).

The stability of the dsDNA form depends not only on the GC-content (% G,C basepairs) but also on sequence (since stacking is sequence specific) and also length (longer molecules are more stable). The stability can be measured in various ways; a common way is the "melting temperature", which is the temperature at which 50% of the ds molecules are converted to ss molecules; melting temperature is dependent on ionic strength and the concentration of DNA. As a result, it is both the percentage of GC base pairs and the overall length of a DNA double helix that determines the strength of the association between the two strands of DNA. Long DNA helices with a high GC-content have stronger-interacting strands, while short helices with high AT content have weaker-interacting strands.[36] In biology, parts of the DNA double helix that need to separate easily, such as the TATAAT Pribnow box in some promoters, tend to have a high AT content, making the strands easier to pull apart.[37]

In the laboratory, the strength of this interaction can be measured by finding the temperature necessary to break the hydrogen bonds, their melting temperature (also called Tm value). When all the base pairs in a DNA double helix melt, the strands separate and exist in solution as two entirely independent molecules. These single-stranded DNA molecules (ssDNA) have no single common shape, but some conformations are more stable than others.[38]

A DNA sequence is called "sense" if its sequence is the same as that of a messenger RNA copy that is translated into protein.[39] The sequence on the opposite strand is called the "antisense" sequence. Both sense and antisense sequences can exist on different parts of the same strand of DNA (i.e. both strands can contain both sense and antisense sequences). In both prokaryotes and eukaryotes, antisense RNA sequences are produced, but the functions of these RNAs are not entirely clear.[40] One proposal is that antisense RNAs are involved in regulating gene expression through RNA-RNA base pairing.[41]

A few DNA sequences in prokaryotes and eukaryotes, and more in plasmids and viruses, blur the distinction between sense and antisense strands by having overlapping genes.[42] In these cases, some DNA sequences do double duty, encoding one protein when read along one strand, and a second protein when read in the opposite direction along the other strand. In bacteria, this overlap may be involved in the regulation of gene transcription,[43] while in viruses, overlapping genes increase the amount of information that can be encoded within the small viral genome.[44]

DNA can be twisted like a rope in a process called DNA supercoiling. With DNA in its "relaxed" state, a strand usually circles the axis of the double helix once every 10.4 base pairs, but if the DNA is twisted the strands become more tightly or more loosely wound.[45] If the DNA is twisted in the direction of the helix, this is positive supercoiling, and the bases are held more tightly together. If they are twisted in the opposite direction, this is negative supercoiling, and the bases come apart more easily. In nature, most DNA has slight negative supercoiling that is introduced by enzymes called topoisomerases.[46] These enzymes are also needed to relieve the twisting stresses introduced into DNA strands during processes such as transcription and DNA replication.[47]

DNA exists in many possible conformations that include A-DNA, B-DNA, and Z-DNA forms, although, only B-DNA and Z-DNA have been directly observed in functional organisms.[17] The conformation that DNA adopts depends on the hydration level, DNA sequence, the amount and direction of supercoiling, chemical modifications of the bases, the type and concentration of metal ions, and the presence of polyamines in solution.[48]

The first published reports of A-DNA X-ray diffraction patternsand also B-DNAused analyses based on Patterson transforms that provided only a limited amount of structural information for oriented fibers of DNA.[49][50] An alternative analysis was then proposed by Wilkins et al., in 1953, for the in vivo B-DNA X-ray diffraction-scattering patterns of highly hydrated DNA fibers in terms of squares of Bessel functions.[51] In the same journal, James Watson and Francis Crick presented their molecular modeling analysis of the DNA X-ray diffraction patterns to suggest that the structure was a double-helix.[11]

Although the B-DNA form is most common under the conditions found in cells,[52] it is not a well-defined conformation but a family of related DNA conformations[53] that occur at the high hydration levels present in living cells. Their corresponding X-ray diffraction and scattering patterns are characteristic of molecular paracrystals with a significant degree of disorder.[54][55]

Compared to B-DNA, the A-DNA form is a wider right-handed spiral, with a shallow, wide minor groove and a narrower, deeper major groove. The A form occurs under non-physiological conditions in partly dehydrated samples of DNA, while in the cell it may be produced in hybrid pairings of DNA and RNA strands, and in enzyme-DNA complexes.[56][57] Segments of DNA where the bases have been chemically modified by methylation may undergo a larger change in conformation and adopt the Z form. Here, the strands turn about the helical axis in a left-handed spiral, the opposite of the more common B form.[58] These unusual structures can be recognized by specific Z-DNA binding proteins and may be involved in the regulation of transcription.[59]

For many years exobiologists have proposed the existence of a shadow biosphere, a postulated microbial biosphere of Earth that uses radically different biochemical and molecular processes than currently known life. One of the proposals was the existence of lifeforms that use arsenic instead of phosphorus in DNA. A report in 2010 of the possibility in the bacterium GFAJ-1, was announced,[60][60][61] though the research was disputed,[61][62] and evidence suggests the bacterium actively prevents the incorporation of arsenic into the DNA backbone and other biomolecules.[63]

At the ends of the linear chromosomes are specialized regions of DNA called telomeres. The main function of these regions is to allow the cell to replicate chromosome ends using the enzyme telomerase, as the enzymes that normally replicate DNA cannot copy the extreme 3 ends of chromosomes.[64] These specialized chromosome caps also help protect the DNA ends, and stop the DNA repair systems in the cell from treating them as damage to be corrected.[65] In human cells, telomeres are usually lengths of single-stranded DNA containing several thousand repeats of a simple TTAGGG sequence.[66]

These guanine-rich sequences may stabilize chromosome ends by forming structures of stacked sets of four-base units, rather than the usual base pairs found in other DNA molecules. Here, four guanine bases form a flat plate and these flat four-base units then stack on top of each other, to form a stable G-quadruplex structure.[68] These structures are stabilized by hydrogen bonding between the edges of the bases and chelation of a metal ion in the centre of each four-base unit.[69] Other structures can also be formed, with the central set of four bases coming from either a single strand folded around the bases, or several different parallel strands, each contributing one base to the central structure.

In addition to these stacked structures, telomeres also form large loop structures called telomere loops, or T-loops. Here, the single-stranded DNA curls around in a long circle stabilized by telomere-binding proteins.[70] At the very end of the T-loop, the single-stranded telomere DNA is held onto a region of double-stranded DNA by the telomere strand disrupting the double-helical DNA and base pairing to one of the two strands. This triple-stranded structure is called a displacement loop or D-loop.[68]

In DNA, fraying occurs when non-complementary regions exist at the end of an otherwise complementary double-strand of DNA. However, branched DNA can occur if a third strand of DNA is introduced and contains adjoining regions able to hybridize with the frayed regions of the pre-existing double-strand. Although the simplest example of branched DNA involves only three strands of DNA, complexes involving additional strands and multiple branches are also possible.[71] Branched DNA can be used in nanotechnology to construct geometric shapes, see the section on uses in technology below.

The expression of genes is influenced by how the DNA is packaged in chromosomes, in a structure called chromatin. Base modifications can be involved in packaging, with regions that have low or no gene expression usually containing high levels of methylation of cytosine bases. DNA packaging and its influence on gene expression can also occur by covalent modifications of the histone protein core around which DNA is wrapped in the chromatin structure or else by remodeling carried out by chromatin remodeling complexes (see Chromatin remodeling). There is, further, crosstalk between DNA methylation and histone modification, so they can coordinately affect chromatin and gene expression.[72]

For one example, cytosine methylation, produces 5-methylcytosine, which is important for X-inactivation of chromosomes.[73] The average level of methylation varies between organisms the worm Caenorhabditis elegans lacks cytosine methylation, while vertebrates have higher levels, with up to 1% of their DNA containing 5-methylcytosine.[74] Despite the importance of 5-methylcytosine, it can deaminate to leave a thymine base, so methylated cytosines are particularly prone to mutations.[75] Other base modifications include adenine methylation in bacteria, the presence of 5-hydroxymethylcytosine in the brain,[76] and the glycosylation of uracil to produce the "J-base" in kinetoplastids.[77][78]

DNA can be damaged by many sorts of mutagens, which change the DNA sequence. Mutagens include oxidizing agents, alkylating agents and also high-energy electromagnetic radiation such as ultraviolet light and X-rays. The type of DNA damage produced depends on the type of mutagen. For example, UV light can damage DNA by producing thymine dimers, which are cross-links between pyrimidine bases.[80] On the other hand, oxidants such as free radicals or hydrogen peroxide produce multiple forms of damage, including base modifications, particularly of guanosine, and double-strand breaks.[81] A typical human cell contains about 150,000 bases that have suffered oxidative damage.[82] Of these oxidative lesions, the most dangerous are double-strand breaks, as these are difficult to repair and can produce point mutations, insertions, deletions from the DNA sequence, and chromosomal translocations.[83] These mutations can cause cancer. Because of inherent limits in the DNA repair mechanisms, if humans lived long enough, they would all eventually develop cancer.[84][85] DNA damages that are naturally occurring, due to normal cellular processes that produce reactive oxygen species, the hydrolytic activities of cellular water, etc., also occur frequently. Although most of these damages are repaired, in any cell some DNA damage may remain despite the action of repair processes. These remaining DNA damages accumulate with age in mammalian postmitotic tissues. This accumulation appears to be an important underlying cause of aging.[86][87][88]

Many mutagens fit into the space between two adjacent base pairs, this is called intercalation. Most intercalators are aromatic and planar molecules; examples include ethidium bromide, acridines, daunomycin, and doxorubicin. For an intercalator to fit between base pairs, the bases must separate, distorting the DNA strands by unwinding of the double helix. This inhibits both transcription and DNA replication, causing toxicity and mutations.[89] As a result, DNA intercalators may be carcinogens, and in the case of thalidomide, a teratogen.[90] Others such as benzo[a]pyrene diol epoxide and aflatoxin form DNA adducts that induce errors in replication.[91] Nevertheless, due to their ability to inhibit DNA transcription and replication, other similar toxins are also used in chemotherapy to inhibit rapidly growing cancer cells.[92]

DNA usually occurs as linear chromosomes in eukaryotes, and circular chromosomes in prokaryotes. The set of chromosomes in a cell makes up its genome; the human genome has approximately 3 billion base pairs of DNA arranged into 46 chromosomes.[93] The information carried by DNA is held in the sequence of pieces of DNA called genes. Transmission of genetic information in genes is achieved via complementary base pairing. For example, in transcription, when a cell uses the information in a gene, the DNA sequence is copied into a complementary RNA sequence through the attraction between the DNA and the correct RNA nucleotides. Usually, this RNA copy is then used to make a matching protein sequence in a process called translation, which depends on the same interaction between RNA nucleotides. In alternative fashion, a cell may simply copy its genetic information in a process called DNA replication. The details of these functions are covered in other articles; here the focus is on the interactions between DNA and other molecules that mediate the function of the genome.

Genomic DNA is tightly and orderly packed in the process called DNA condensation, to fit the small available volumes of the cell. In eukaryotes, DNA is located in the cell nucleus, with small amounts in mitochondria and chloroplasts. In prokaryotes, the DNA is held within an irregularly shaped body in the cytoplasm called the nucleoid.[94] The genetic information in a genome is held within genes, and the complete set of this information in an organism is called its genotype. A gene is a unit of heredity and is a region of DNA that influences a particular characteristic in an organism. Genes contain an open reading frame that can be transcribed, and regulatory sequences such as promoters and enhancers, which control transcription of the open reading frame.

In many species, only a small fraction of the total sequence of the genome encodes protein. For example, only about 1.5% of the human genome consists of protein-coding exons, with over 50% of human DNA consisting of non-coding repetitive sequences.[95] The reasons for the presence of so much noncoding DNA in eukaryotic genomes and the extraordinary differences in genome size, or C-value, among species represent a long-standing puzzle known as the "C-value enigma".[96] However, some DNA sequences that do not code protein may still encode functional non-coding RNA molecules, which are involved in the regulation of gene expression.[97]

Some noncoding DNA sequences play structural roles in chromosomes. Telomeres and centromeres typically contain few genes, but are important for the function and stability of chromosomes.[65][99] An abundant form of noncoding DNA in humans are pseudogenes, which are copies of genes that have been disabled by mutation.[100] These sequences are usually just molecular fossils, although they can occasionally serve as raw genetic material for the creation of new genes through the process of gene duplication and divergence.[101]

A gene is a sequence of DNA that contains genetic information and can influence the phenotype of an organism. Within a gene, the sequence of bases along a DNA strand defines a messenger RNA sequence, which then defines one or more protein sequences. The relationship between the nucleotide sequences of genes and the amino-acid sequences of proteins is determined by the rules of translation, known collectively as the genetic code. The genetic code consists of three-letter 'words' called codons formed from a sequence of three nucleotides (e.g. ACT, CAG, TTT).

In transcription, the codons of a gene are copied into messenger RNA by RNA polymerase. This RNA copy is then decoded by a ribosome that reads the RNA sequence by base-pairing the messenger RNA to transfer RNA, which carries amino acids. Since there are 4 bases in 3-letter combinations, there are 64 possible codons (43combinations). These encode the twenty standard amino acids, giving most amino acids more than one possible codon. There are also three 'stop' or 'nonsense' codons signifying the end of the coding region; these are the TAA, TGA, and TAG codons.

Cell division is essential for an organism to grow, but, when a cell divides, it must replicate the DNA in its genome so that the two daughter cells have the same genetic information as their parent. The double-stranded structure of DNA provides a simple mechanism for DNA replication. Here, the two strands are separated and then each strand's complementary DNA sequence is recreated by an enzyme called DNA polymerase. This enzyme makes the complementary strand by finding the correct base through complementary base pairing, and bonding it onto the original strand. As DNA polymerases can only extend a DNA strand in a 5 to 3 direction, different mechanisms are used to copy the antiparallel strands of the double helix.[102] In this way, the base on the old strand dictates which base appears on the new strand, and the cell ends up with a perfect copy of its DNA.

Naked extracellular DNA (eDNA), most of it released by cell death, is nearly ubiquitous in the environment. Its concentration in soil may be as high as 2 g/L, and its concentration in natural aquatic environments may be as high at 88 g/L.[103] Various possible functions have been proposed for eDNA: it may be involved in horizontal gene transfer;[104] it may provide nutrients;[105] and it may act as a buffer to recruit or titrate ions or antibiotics.[106] Extracellular DNA acts as a functional extracellular matrix component in the biofilms of several bacterial species. It may act as a recognition factor to regulate the attachment and dispersal of specific cell types in the biofilm;[107] it may contribute to biofilm formation;[108] and it may contribute to the biofilm's physical strength and resistance to biological stress.[109]

All the functions of DNA depend on interactions with proteins. These protein interactions can be non-specific, or the protein can bind specifically to a single DNA sequence. Enzymes can also bind to DNA and of these, the polymerases that copy the DNA base sequence in transcription and DNA replication are particularly important.

Structural proteins that bind DNA are well-understood examples of non-specific DNA-protein interactions. Within chromosomes, DNA is held in complexes with structural proteins. These proteins organize the DNA into a compact structure called chromatin. In eukaryotes this structure involves DNA binding to a complex of small basic proteins called histones, while in prokaryotes multiple types of proteins are involved.[110][111] The histones form a disk-shaped complex called a nucleosome, which contains two complete turns of double-stranded DNA wrapped around its surface. These non-specific interactions are formed through basic residues in the histones, making ionic bonds to the acidic sugar-phosphate backbone of the DNA, and are thus largely independent of the base sequence.[112] Chemical modifications of these basic amino acid residues include methylation, phosphorylation and acetylation.[113] These chemical changes alter the strength of the interaction between the DNA and the histones, making the DNA more or less accessible to transcription factors and changing the rate of transcription.[114] Other non-specific DNA-binding proteins in chromatin include the high-mobility group proteins, which bind to bent or distorted DNA.[115] These proteins are important in bending arrays of nucleosomes and arranging them into the larger structures that make up chromosomes.[116]

A distinct group of DNA-binding proteins are the DNA-binding proteins that specifically bind single-stranded DNA. In humans, replication protein A is the best-understood member of this family and is used in processes where the double helix is separated, including DNA replication, recombination and DNA repair.[117] These binding proteins seem to stabilize single-stranded DNA and protect it from forming stem-loops or being degraded by nucleases.

In contrast, other proteins have evolved to bind to particular DNA sequences. The most intensively studied of these are the various transcription factors, which are proteins that regulate transcription. Each transcription factor binds to one particular set of DNA sequences and activates or inhibits the transcription of genes that have these sequences close to their promoters. The transcription factors do this in two ways. Firstly, they can bind the RNA polymerase responsible for transcription, either directly or through other mediator proteins; this locates the polymerase at the promoter and allows it to begin transcription.[119] Alternatively, transcription factors can bind enzymes that modify the histones at the promoter. This changes the accessibility of the DNA template to the polymerase.[120]

As these DNA targets can occur throughout an organism's genome, changes in the activity of one type of transcription factor can affect thousands of genes.[121] Consequently, these proteins are often the targets of the signal transduction processes that control responses to environmental changes or cellular differentiation and development. The specificity of these transcription factors' interactions with DNA come from the proteins making multiple contacts to the edges of the DNA bases, allowing them to "read" the DNA sequence. Most of these base-interactions are made in the major groove, where the bases are most accessible.[34]

Nucleases are enzymes that cut DNA strands by catalyzing the hydrolysis of the phosphodiester bonds. Nucleases that hydrolyse nucleotides from the ends of DNA strands are called exonucleases, while endonucleases cut within strands. The most frequently used nucleases in molecular biology are the restriction endonucleases, which cut DNA at specific sequences. For instance, the EcoRV enzyme shown to the left recognizes the 6-base sequence 5-GATATC-3 and makes a cut at the horizontal line. In nature, these enzymes protect bacteria against phage infection by digesting the phage DNA when it enters the bacterial cell, acting as part of the restriction modification system.[123] In technology, these sequence-specific nucleases are used in molecular cloning and DNA fingerprinting.

Enzymes called DNA ligases can rejoin cut or broken DNA strands.[124] Ligases are particularly important in lagging strand DNA replication, as they join together the short segments of DNA produced at the replication fork into a complete copy of the DNA template. They are also used in DNA repair and genetic recombination.[124]

Topoisomerases are enzymes with both nuclease and ligase activity. These proteins change the amount of supercoiling in DNA. Some of these enzymes work by cutting the DNA helix and allowing one section to rotate, thereby reducing its level of supercoiling; the enzyme then seals the DNA break.[46] Other types of these enzymes are capable of cutting one DNA helix and then passing a second strand of DNA through this break, before rejoining the helix.[125] Topoisomerases are required for many processes involving DNA, such as DNA replication and transcription.[47]

Helicases are proteins that are a type of molecular motor. They use the chemical energy in nucleoside triphosphates, predominantly adenosine triphosphate (ATP), to break hydrogen bonds between bases and unwind the DNA double helix into single strands.[126] These enzymes are essential for most processes where enzymes need to access the DNA bases.

Polymerases are enzymes that synthesize polynucleotide chains from nucleoside triphosphates. The sequence of their products are created based on existing polynucleotide chainswhich are called templates. These enzymes function by repeatedly adding a nucleotide to the 3 hydroxyl group at the end of the growing polynucleotide chain. As a consequence, all polymerases work in a 5 to 3 direction.[127] In the active site of these enzymes, the incoming nucleoside triphosphate base-pairs to the template: this allows polymerases to accurately synthesize the complementary strand of their template. Polymerases are classified according to the type of template that they use.

In DNA replication, DNA-dependent DNA polymerases make copies of DNA polynucleotide chains. To preserve biological information, it is essential that the sequence of bases in each copy are precisely complementary to the sequence of bases in the template strand. Many DNA polymerases have a proofreading activity. Here, the polymerase recognizes the occasional mistakes in the synthesis reaction by the lack of base pairing between the mismatched nucleotides. If a mismatch is detected, a 3 to 5 exonuclease activity is activated and the incorrect base removed.[128] In most organisms, DNA polymerases function in a large complex called the replisome that contains multiple accessory subunits, such as the DNA clamp or helicases.[129]

RNA-dependent DNA polymerases are a specialized class of polymerases that copy the sequence of an RNA strand into DNA. They include reverse transcriptase, which is a viral enzyme involved in the infection of cells by retroviruses, and telomerase, which is required for the replication of telomeres.[64][130] Telomerase is an unusual polymerase because it contains its own RNA template as part of its structure.[65]

Transcription is carried out by a DNA-dependent RNA polymerase that copies the sequence of a DNA strand into RNA. To begin transcribing a gene, the RNA polymerase binds to a sequence of DNA called a promoter and separates the DNA strands. It then copies the gene sequence into a messenger RNA transcript until it reaches a region of DNA called the terminator, where it halts and detaches from the DNA. As with human DNA-dependent DNA polymerases, RNA polymerase II, the enzyme that transcribes most of the genes in the human genome, operates as part of a large protein complex with multiple regulatory and accessory subunits.[131]

A DNA helix usually does not interact with other segments of DNA, and in human cells the different chromosomes even occupy separate areas in the nucleus called "chromosome territories".[133] This physical separation of different chromosomes is important for the ability of DNA to function as a stable repository for information, as one of the few times chromosomes interact is in chromosomal crossover which occurs during sexual reproduction, when genetic recombination occurs. Chromosomal crossover is when two DNA helices break, swap a section and then rejoin.

Recombination allows chromosomes to exchange genetic information and produces new combinations of genes, which increases the efficiency of natural selection and can be important in the rapid evolution of new proteins.[134] Genetic recombination can also be involved in DNA repair, particularly in the cell's response to double-strand breaks.[135]

The most common form of chromosomal crossover is homologous recombination, where the two chromosomes involved share very similar sequences. Non-homologous recombination can be damaging to cells, as it can produce chromosomal translocations and genetic abnormalities. The recombination reaction is catalyzed by enzymes known as recombinases, such as RAD51.[136] The first step in recombination is a double-stranded break caused by either an endonuclease or damage to the DNA.[137] A series of steps catalyzed in part by the recombinase then leads to joining of the two helices by at least one Holliday junction, in which a segment of a single strand in each helix is annealed to the complementary strand in the other helix. The Holliday junction is a tetrahedral junction structure that can be moved along the pair of chromosomes, swapping one strand for another. The recombination reaction is then halted by cleavage of the junction and re-ligation of the released DNA.[138]

DNA contains the genetic information that allows all modern living things to function, grow and reproduce. However, it is unclear how long in the 4-billion-year history of life DNA has performed this function, as it has been proposed that the earliest forms of life may have used RNA as their genetic material.[139][140] RNA may have acted as the central part of early cell metabolism as it can both transmit genetic information and carry out catalysis as part of ribozymes.[141] This ancient RNA world where nucleic acid would have been used for both catalysis and genetics may have influenced the evolution of the current genetic code based on four nucleotide bases. This would occur, since the number of different bases in such an organism is a trade-off between a small number of bases increasing replication accuracy and a large number of bases increasing the catalytic efficiency of ribozymes.[142] However, there is no direct evidence of ancient genetic systems, as recovery of DNA from most fossils is impossible because DNA survives in the environment for less than one million years, and slowly degrades into short fragments in solution.[143] Claims for older DNA have been made, most notably a report of the isolation of a viable bacterium from a salt crystal 250 million years old,[144] but these claims are controversial.[145][146]

Building blocks of DNA (adenine, guanine and related organic molecules) may have been formed extraterrestrially in outer space.[147][148][149] Complex DNA and RNA organic compounds of life, including uracil, cytosine, and thymine, have also been formed in the laboratory under conditions mimicking those found in outer space, using starting chemicals, such as pyrimidine, found in meteorites. Pyrimidine, like polycyclic aromatic hydrocarbons (PAHs), the most carbon-rich chemical found in the universe, may have been formed in red giants or in interstellar cosmic dust and gas clouds.[150]

Methods have been developed to purify DNA from organisms, such as phenol-chloroform extraction, and to manipulate it in the laboratory, such as restriction digests and the polymerase chain reaction. Modern biology and biochemistry make intensive use of these techniques in recombinant DNA technology. Recombinant DNA is a man-made DNA sequence that has been assembled from other DNA sequences. They can be transformed into organisms in the form of plasmids or in the appropriate format, by using a viral vector.[151] The genetically modified organisms produced can be used to produce products such as recombinant proteins, used in medical research,[152] or be grown in agriculture.[153][154]

Forensic scientists can use DNA in blood, semen, skin, saliva or hair found at a crime scene to identify a matching DNA of an individual, such as a perpetrator. This process is formally termed DNA profiling, but may also be called "genetic fingerprinting". In DNA profiling, the lengths of variable sections of repetitive DNA, such as short tandem repeats and minisatellites, are compared between people. This method is usually an extremely reliable technique for identifying a matching DNA.[155] However, identification can be complicated if the scene is contaminated with DNA from several people.[156] DNA profiling was developed in 1984 by British geneticist Sir Alec Jeffreys,[157] and first used in forensic science to convict Colin Pitchfork in the 1988 Enderby murders case.[158]

The development of forensic science, and the ability to now obtain genetic matching on minute samples of blood, skin, saliva, or hair has led to re-examining many cases. Evidence can now be uncovered that was scientifically impossible at the time of the original examination. Combined with the removal of the double jeopardy law in some places, this can allow cases to be reopened where prior trials have failed to produce sufficient evidence to convince a jury. People charged with serious crimes may be required to provide a sample of DNA for matching purposes. The most obvious defence to DNA matches obtained forensically is to claim that cross-contamination of evidence has occurred. This has resulted in meticulous strict handling procedures with new cases of serious crime. DNA profiling is also used successfully to positively identify victims of mass casualty incidents,[159] bodies or body parts in serious accidents, and individual victims in mass war graves, via matching to family members.

DNA profiling is also used in DNA paternity testing to determine if someone is the biological parent or grandparent of a child with the probability of parentage is typically 99.99% when the alleged parent is biologically related to the child. Normal DNA sequencing methods happen after birth but there are new methods to test paternity while a mother is still pregnant.[160]

Deoxyribozymes, also called DNAzymes or catalytic DNA are first discovered in 1994.[161] They are mostly single stranded DNA sequences isolated from a large pool of random DNA sequences through a combinatorial approach called in vitro selection or systematic evolution of ligands by exponential enrichment (SELEX). DNAzymes catalyze variety of chemical reactions including RNA-DNA cleavage, RNA-DNA ligation, amino acids phosphorylation-dephosphorylation, carbon-carbon bond formation, and etc. DNAzymes can enhance catalytic rate of chemical reactions up to 100,000,000,000-fold over the uncatalyzed reaction.[162] The most extensively studied class of DNAzymes are RNA-cleaving types which have been used to detect different metal ions and designing therapeutic agents. Several metal-specific DNAzymes have been reported including the GR-5 DNAzyme (lead-specific),[161] the CA1-3 DNAzymes (copper-specific),[163] the 39E DNAzyme (uranyl-specific) and the NaA43 DNAzyme (sodium-specific).[164] The NaA43 DNAzyme, which is reported to be more than 10,000-fold selective for sodium over other metal ions, was used to make a real-time sodium sensor in living cells.

Bioinformatics involves the development of techniques to store, data mine, search and manipulate biological data, including DNA nucleic acid sequence data. These have led to widely applied advances in computer science, especially string searching algorithms, machine learning and database theory.[165] String searching or matching algorithms, which find an occurrence of a sequence of letters inside a larger sequence of letters, were developed to search for specific sequences of nucleotides.[166] The DNA sequence may be aligned with other DNA sequences to identify homologous sequences and locate the specific mutations that make them distinct. These techniques, especially multiple sequence alignment, are used in studying phylogenetic relationships and protein function.[167] Data sets representing entire genomes' worth of DNA sequences, such as those produced by the Human Genome Project, are difficult to use without the annotations that identify the locations of genes and regulatory elements on each chromosome. Regions of DNA sequence that have the characteristic patterns associated with protein- or RNA-coding genes can be identified by gene finding algorithms, which allow researchers to predict the presence of particular gene products and their possible functions in an organism even before they have been isolated experimentally.[168] Entire genomes may also be compared, which can shed light on the evolutionary history of particular organism and permit the examination of complex evolutionary events.

DNA nanotechnology uses the unique molecular recognition properties of DNA and other nucleic acids to create self-assembling branched DNA complexes with useful properties.[169] DNA is thus used as a structural material rather than as a carrier of biological information. This has led to the creation of two-dimensional periodic lattices (both tile-based and using the DNA origami method) and three-dimensional structures in the shapes of polyhedra.[170]Nanomechanical devices and algorithmic self-assembly have also been demonstrated,[171] and these DNA structures have been used to template the arrangement of other molecules such as gold nanoparticles and streptavidin proteins.[172]

Because DNA collects mutations over time, which are then inherited, it contains historical information, and, by comparing DNA sequences, geneticists can infer the evolutionary history of organisms, their phylogeny.[173] This field of phylogenetics is a powerful tool in evolutionary biology. If DNA sequences within a species are compared, population geneticists can learn the history of particular populations. This can be used in studies ranging from ecological genetics to anthropology; For example, DNA evidence is being used to try to identify the Ten Lost Tribes of Israel.[174][175]

In a paper published in Nature in January 2013, scientists from the European Bioinformatics Institute and Agilent Technologies proposed a mechanism to use DNA's ability to code information as a means of digital data storage. The group was able to encode 739 kilobytes of data into DNA code, synthesize the actual DNA, then sequence the DNA and decode the information back to its original form, with a reported 100% accuracy. The encoded information consisted of text files and audio files. A prior experiment was published in August 2012. It was conducted by researchers at Harvard University, where the text of a 54,000-word book was encoded in DNA.[176][177]

DNA was first isolated by the Swiss physician Friedrich Miescher who, in 1869, discovered a microscopic substance in the pus of discarded surgical bandages. As it resided in the nuclei of cells, he called it "nuclein".[178][179] In 1878, Albrecht Kossel isolated the non-protein component of "nuclein", nucleic acid, and later isolated its five primary nucleobases.[180][181] In 1919, Phoebus Levene identified the base, sugar and phosphate nucleotide unit.[182] Levene suggested that DNA consisted of a string of nucleotide units linked together through the phosphate groups. Levene thought the chain was short and the bases repeated in a fixed order. In 1937, William Astbury produced the first X-ray diffraction patterns that showed that DNA had a regular structure.[183]

In 1927, Nikolai Koltsov proposed that inherited traits would be inherited via a "giant hereditary molecule" made up of "two mirror strands that would replicate in a semi-conservative fashion using each strand as a template".[184][185] In 1928, Frederick Griffith in his experiment discovered that traits of the "smooth" form of Pneumococcus could be transferred to the "rough" form of the same bacteria by mixing killed "smooth" bacteria with the live "rough" form.[186][187] This system provided the first clear suggestion that DNA carries genetic informationthe AveryMacLeodMcCarty experimentwhen Oswald Avery, along with coworkers Colin MacLeod and Maclyn McCarty, identified DNA as the transforming principle in 1943.[188] DNA's role in heredity was confirmed in 1952, when Alfred Hershey and Martha Chase in the HersheyChase experiment showed that DNA is the genetic material of the T2 phage.[189]

In 1953, James Watson and Francis Crick suggested what is now accepted as the first correct double-helix model of DNA structure in the journal Nature.[11] Their double-helix, molecular model of DNA was then based on one X-ray diffraction image (labeled as "Photo 51")[190] taken by Rosalind Franklin and Raymond Gosling in May 1952, and the information that the DNA bases are paired.

Experimental evidence supporting the Watson and Crick model was published in a series of five articles in the same issue of Nature.[191] Of these, Franklin and Gosling's paper was the first publication of their own X-ray diffraction data and original analysis method that partly supported the Watson and Crick model;[50][192] this issue also contained an article on DNA structure by Maurice Wilkins and two of his colleagues, whose analysis and in vivo B-DNA X-ray patterns also supported the presence in vivo of the double-helical DNA configurations as proposed by Crick and Watson for their double-helix molecular model of DNA in the prior two pages of Nature.[51] In 1962, after Franklin's death, Watson, Crick, and Wilkins jointly received the Nobel Prize in Physiology or Medicine.[193] Nobel Prizes are awarded only to living recipients. A debate continues about who should receive credit for the discovery.[194]

In an influential presentation in 1957, Crick laid out the central dogma of molecular biology, which foretold the relationship between DNA, RNA, and proteins, and articulated the "adaptor hypothesis".[195] Final confirmation of the replication mechanism that was implied by the double-helical structure followed in 1958 through the MeselsonStahl experiment.[196] Further work by Crick and coworkers showed that the genetic code was based on non-overlapping triplets of bases, called codons, allowing Har Gobind Khorana, Robert W. Holley and Marshall Warren Nirenberg to decipher the genetic code.[197] These findings represent the birth of molecular biology.

Read more from the original source:
DNA - Wikipedia

Read More...

BRCA1 and BRCA2: Cancer Risk and Genetic Testing Fact …

Wednesday, October 5th, 2016

What are BRCA1 and BRCA2?

BRCA1 and BRCA2 are human genes that produce tumor suppressor proteins. These proteins help repair damaged DNA and, therefore, play a role in ensuring the stability of the cells genetic material. When either of these genes is mutated, or altered, such that its protein product either is not made or does not function correctly, DNA damage may not be repaired properly. As a result, cells are more likely to develop additional genetic alterations that can lead to cancer.

Specific inherited mutations in BRCA1 and BRCA2 increase the risk of female breast and ovarian cancers, and they have been associated with increased risks of several additional types of cancer. Together, BRCA1 and BRCA2 mutations account for about 20 to 25 percent of hereditary breast cancers (1) and about 5 to 10 percent of all breast cancers (2). In addition, mutations in BRCA1 and BRCA2 account for around 15 percent of ovarian cancers overall (3). Breast and ovarian cancers associated with BRCA1 and BRCA2 mutations tend to develop at younger ages than their nonhereditary counterparts.

A harmful BRCA1 or BRCA2 mutation can be inherited from a persons mother or father. Each child of a parent who carries a mutation in one of these genes has a 50 percent chance (or 1 chance in 2) of inheriting the mutation. The effects of mutations in BRCA1 and BRCA2 are seen even when a persons second copy of the gene is normal.

How much does having a BRCA1 or BRCA2 gene mutation increase a womans risk of breast and ovarian cancer?

A womans lifetime risk of developing breast and/or ovarian cancer is greatly increased if she inherits a harmful mutation in BRCA1 or BRCA2.

Breast cancer: About 12 percent of women in the general population will develop breast cancer sometime during their lives (4). By contrast, according to the most recent estimates, 55 to 65 percent of women who inherit a harmful BRCA1 mutation and around 45 percent of women who inherit a harmful BRCA2 mutation will develop breast cancer by age 70 years (5, 6).

Ovarian cancer: About 1.3 percent of women in the general population will develop ovarian cancer sometime during their lives (4). By contrast, according to the most recent estimates, 39 percent of women who inherit a harmful BRCA1 mutation (5, 6) and 11 to 17 percent of women who inherit a harmful BRCA2 mutation will develop ovarian cancer by age 70 years (5, 6).

It is important to note that these estimated percentages of lifetime risk are different from those available previously; the estimates have changed as more information has become available, and they may change again with additional research. No long-term general population studies have directly compared cancer risk in women who have and do not have a harmful BRCA1 or BRCA2 mutation.

It is also important to note that other characteristics of a particular woman can make her cancer risk higher or lower than the average risks. These characteristics include her family history ofbreast, ovarian, and, possibly, other cancers; the specific mutation(s) she has inherited; and other risk factors, suchas her reproductivehistory. However, at this time, based on current data, none of these other factors seems to be as strong as the effect of carrying a harmful BRCA1 or BRCA2 mutation.

What other cancers have been linked to mutations in BRCA1 and BRCA2?

Are mutations in BRCA1 and BRCA2 more common in certain racial/ethnic populations than others?

Yes. For example, people of Ashkenazi Jewish descent have a higher prevalence of harmful BRCA1 and BRCA2 mutations than people in the general U.S. population. Other ethnic and geographic populations around the world, such as the Norwegian, Dutch, and Icelandic peoples, also have a higher prevalence of specific harmful BRCA1 and BRCA2 mutations.

In addition, limited data indicate that the prevalence of specific harmful BRCA1 and BRCA2 mutations may vary among individual racial and ethnic groups in the United States, including African Americans, Hispanics, Asian Americans, and non-Hispanic whites (15, 16).

Are genetic tests available to detect BRCA1 and BRCA2 mutations?

Yes. Several different tests are available, including tests that look for a known mutation in one of the genes (i.e., a mutation that has already been identified in another family member) and tests that check for all possible mutations in both genes. DNA (from a blood or saliva sample) is needed for mutation testing. The sample is sent to a laboratory for analysis. It usually takes about a month to get the test results.

Who should consider genetic testing for BRCA1 and BRCA2 mutations?

Because harmful BRCA1 and BRCA2 gene mutations are relatively rare in the general population, most experts agree that mutation testing of individuals who do not have cancer should be performed only when the persons individual or family history suggests the possible presence of a harmful mutation in BRCA1 or BRCA2.

In December 2013, the United States Preventive Services Task Force recommended that women who have family members with breast, ovarian, fallopian tube, or peritoneal cancer be evaluated to see if they have a family history that is associated with an increased risk of a harmful mutation in one of these genes (17).

Several screening tools are now available to help health care providers with this evaluation (17). These tools assess family history factors that are associated with an increased likelihood of having a harmful mutation in BRCA1 or BRCA2, including:

When an individual has a family history that is suggestive of the presence of a BRCA1 or BRCA2 mutation, it may be most informative to first test a family member who has cancer if that person is still alive and willing to be tested. If that person is found to have a harmful BRCA1 or BRCA2 mutation, then other family members may want to consider genetic counseling to learn more about their potential risks and whether genetic testing for mutations in BRCA1 and BRCA2 might be appropriate for them.

If it is not possible to confirm the presence of a harmful BRCA1 or BRCA2 mutation in a family member who has cancer, it is appropriate for both men and women who do not have cancer but have a family medical history that suggests the presence of such a mutation to have genetic counseling for possible testing.

Some individualsfor example, those who were adopted at birthmay not know their family history. In cases where a woman with an unknown family history has an early-onset breast cancer or ovarian cancer or a man with an unknown family history is diagnosed with breast cancer, it may be reasonable for that individual to consider genetic testing for a BRCA1 or BRCA2 mutation. Individuals with an unknown family history who do not have an early-onset cancer or male breast cancer are at very low risk of having a harmful BRCA1 or BRCA2 mutation and are unlikely to benefit from routine genetic testing.

Professional societies do not recommend that children, even those with a family history suggestive of a harmful BRCA1 or BRCA2 mutation, undergo genetic testing for BRCA1 or BRCA2. This is because no risk-reduction strategies exist for children, and children's risks of developing a cancer type associated with a BRCA1 or BRCA2 mutation are extremely low. After children with a family history suggestive of a harmful BRCA1 or BRCA2 mutation become adults, however, they may want to obtain genetic counseling about whether or not to undergoing genetic testing.

Should people considering genetic testing for BRCA1 and BRCA2 mutations talk with a genetic counselor?

Genetic counseling is generally recommended before and after any genetic test for an inherited cancer syndrome. This counseling should be performed by a health care professional who is experienced in cancer genetics. Genetic counseling usually covers many aspects of the testing process, including:

How much does BRCA1 and BRCA2 mutation testing cost?

The Affordable Care Act considers genetic counseling and BRCA1 and BRCA2 mutation testing for individuals at high risk a covered preventive service. People considering BRCA1 and BRCA2 mutation testing may want to confirm their insurance coverage for genetic tests before having the test.

Some of the genetic testing companies that offer testing for BRCA1 and BRCA2 mutations may offer testing at no charge to patients who lack insurance and meet specific financial and medical criteria.

What does a positive BRCA1 or BRCA2 genetic test result mean?

BRCA1 and BRCA2 gene mutation testing can give several possible results: a positive result, a negative result, or an ambiguous or uncertain result.

A positive test result indicates that a person has inherited a known harmful mutation in BRCA1 or BRCA2 and, therefore, has an increased risk of developing certain cancers. However, a positive test result cannot tell whether or when an individual will actually develop cancer. For example, some women who inherit a harmful BRCA1 or BRCA2 mutation will never develop breast or ovarian cancer.

A positive genetic test result may also have important health and social implications for family members, including future generations. Unlike most other medical tests, genetic tests can reveal information not only about the person being tested but also about that persons relatives:

What does a negative BRCA1 or BRCA2 test result mean?

A negative test result can be more difficult to understand than a positive result because what the result means depends in part on an individuals family history of cancer and whether a BRCA1 or BRCA2 mutation has been identified in a blood relative.

If a close (first- or second-degree) relative of the tested person is known to carry a harmful BRCA1 or BRCA2 mutation, a negative test result is clear: it means that person does not carry the harmful mutation that is responsible for the familial cancer, and thus cannot pass it on to their children. Such a test result is called a true negative. A person with such a test result is currently thought to have the same risk of cancer as someone in the general population.

If the tested person has a family history that suggests the possibility of having a harmful mutation in BRCA1 or BRCA2 but complete gene testing identifies no such mutation in the family, a negative result is less clear. The likelihood that genetic testing will miss a known harmful BRCA1 or BRCA2 mutation is very low, but it could happen. Moreover, scientists continue to discover new BRCA1 and BRCA2 mutations and have not yet identified all potentially harmful ones. Therefore, it is possible that a person in this scenario with a "negative" test result actually has an as-yet unknown harmful BRCA1 or BRCA2 mutation that has not been identified.

It is also possible for people to have a mutation in a gene other than BRCA1 or BRCA2 that increases their cancer risk but is not detectable by the test used. People considering genetic testing for BRCA1 and BRCA2 mutations may want to discuss these potential uncertainties with a genetic counselor before undergoing testing.

What does an ambiguous or uncertain BRCA1 or BRCA2 test result mean?

Sometimes, a genetic test finds a change in BRCA1 or BRCA2 that has not been previously associated with cancer. This type of test result may be described as ambiguous (often referred to as a genetic variant of uncertain significance) because it isnt known whether this specific gene change affects a persons risk of developing cancer. One study found that 10 percent of women who underwent BRCA1 and BRCA2 mutation testing had this type of ambiguous result (18).

As more research is conducted and more people are tested for BRCA1 and BRCA2 mutations, scientists will learn more about these changes and cancer risk. Genetic counseling can help a person understand what an ambiguous change in BRCA1 or BRCA2 may mean in terms of cancer risk. Over time, additional studies of variants of uncertain significance may result in a specific mutation being re-classified as either harmful or clearly not harmful.

How can a person who has a positive test result manage their risk of cancer?

Several options are available for managing cancer risk in individuals who have a known harmful BRCA1 or BRCA2 mutation. These include enhanced screening, prophylactic (risk-reducing) surgery, and chemoprevention.

Enhanced Screening. Some women who test positive for BRCA1 and BRCA2 mutations may choose to start cancer screening at younger ages than the general population or to have more frequent screening. For example, some experts recommend that women who carry a harmful BRCA1 or BRCA2 mutation undergo clinical breast examinations beginning at age 25 to 35 years (19). And some expert groups recommend that women who carry such a mutation have a mammogram every year, beginning at age 25 to 35 years.

Enhanced screening may increase the chance of detecting breast cancer at an early stage, when it may have a better chance of being treated successfully. Women who have a positive test result should ask their health care provider about the possible harms of diagnostic tests that involve radiation (mammograms or x-rays).

Recent studies have shown that MRI may be more sensitive than mammography for women at high risk of breast cancer (20, 21). However, mammography can also identify some breast cancers that are not identified by MRI (22), and MRI may be less specific (i.e., lead to more false-positive results) than mammography. Several organizations, such as the American Cancer Society and the National Comprehensive Cancer Network, now recommend annual screening with mammography and MRI for women who have a high risk of breast cancer.

No effective ovarian cancer screening methods currently exist. Some groups recommend transvaginal ultrasound, blood tests for the antigen CA-125, and clinical examinations for ovarian cancer screening in women with harmful BRCA1 or BRCA2 mutations, but none of these methods appears to detect ovarian tumors at an early enough stage to reduce the risk of dying from ovarian cancer (23). For a screening method to be considered effective, it must have demonstrated reduced mortality from the disease of interest. This standard has not yet been met for ovarian cancer screening.

The benefits of screening for breast and other cancers in men who carry harmful mutations in BRCA1 or BRCA2 is also not known, but some expert groups recommend that men who are known to carry a harmful mutation undergo regular mammography as well as testing for prostate cancer. The value of these screening strategies remains unproven at present.

Prophylactic (Risk-reducing) Surgery. Prophylactic surgery involves removing as much of the "at-risk" tissue as possible. Women may choose to have both breasts removed (bilateral prophylactic mastectomy) to reduce their risk of breast cancer. Surgery to remove a woman's ovaries and fallopian tubes (bilateral prophylactic salpingo-oophorectomy) can help reduce her risk of ovarian cancer. Removing the ovaries also reduces the risk of breast cancer in premenopausal women by eliminating a source of hormones that can fuel the growth of some types of breast cancer.

No evidence is available regarding the effectiveness of bilateral prophylactic mastectomy in reducing breast cancer risk in men with a harmful BRCA1 or BRCA2 mutation or a family history of breast cancer. Therefore, bilateral prophylactic mastectomy for men at high risk of breast cancer is considered an experimental procedure, and insurance companies will not normally cover it.

Prophylactic surgery does not completely guarantee that cancer will not develop because not all at-risk tissue can be removed by these procedures. Some women have developed breast cancer, ovarian cancer, or primary peritoneal carcinomatosis (a type of cancer similar to ovarian cancer) even after prophylactic surgery. Nevertheless, the mortality reduction associated with this surgery is substantial: Research demonstrates that women who underwent bilateral prophylactic salpingo-oophorectomy had a nearly 80 percent reduction in risk of dying from ovarian cancer, a 56 percent reduction in risk of dying from breast cancer (24), and a 77 percent reduction in risk of dying from any cause (25).

Emerging evidence (25) suggests that the amount of protection that removing the ovaries and fallopian tubes provides against the development of breast and ovarian cancer may be similar for carriers of BRCA1 and BRCA2 mutations, in contrast to earlier studies (26).

Chemoprevention. Chemoprevention is the use of drugs, vitamins, or other agents to try to reduce the risk of, or delay the recurrence of, cancer. Although two chemopreventive drugs (tamoxifen and raloxifene) have been approved by the U.S. Food and Drug Administration (FDA) to reduce the risk of breast cancer in women at increased risk, the role of these drugs in women with harmful BRCA1 or BRCA2 mutations is not yet clear.

Data from three studies suggest that tamoxifen may be able to help lower the risk of breast cancer in BRCA1 and BRCA2 mutation carriers (27), including the risk of cancer in the opposite breast among women previously diagnosed with breast cancer (28, 29). Studies have not examined the effectiveness of raloxifene in BRCA1 and BRCA2 mutation carriers specifically.

Oral contraceptives (birth control pills) are thought to reduce the risk of ovarian cancer by about 50 percent both in the general population and in women with harmful BRCA1 or BRCA2 mutations (30).

What are some of the benefits of genetic testing for breast and ovarian cancer risk?

There can be benefits to genetic testing, regardless of whether a person receives a positive or a negative result.

The potential benefits of a true negative result include a sense of relief regarding the future risk of cancer, learning that one's children are not at risk of inheriting the family's cancer susceptibility, and the possibility that special checkups, tests, or preventive surgeries may not be needed.

A positive test result may bring relief by resolving uncertainty regarding future cancer risk and may allow people to make informed decisions about their future, including taking steps to reduce their cancer risk. In addition, people who have a positive test result may choose to participate in medical research that could, in the long run, help reduce deaths from hereditary breast and ovarian cancer.

What are some of the possible harms of genetic testing for breast and ovarian cancer risk?

The direct medical harms of genetic testing are minimal, but knowledge of test results may have harmful effects on a persons emotions, social relationships, finances, and medical choices.

People who receive a positive test result may feel anxious, depressed, or angry. They may have difficulty making choices about whether to have preventive surgery or about which surgery to have.

People who receive a negative test result may experience survivor guilt, caused by the knowledge that they likely do not have an increased risk of developing a disease that affects one or more loved ones.

Because genetic testing can reveal information about more than one family member, the emotions caused by test results can create tension within families. Test results can also affect personal life choices, such as decisions about career, marriage, and childbearing.

Violations of privacy and of the confidentiality of genetic test results are additional potential risks. However, the federal Health Insurance Portability and Accountability Act and various state laws protect the privacy of a persons genetic information. Moreover, the federal Genetic Information Nondiscrimination Act, along with many state laws, prohibits discrimination based on genetic information in relation to health insurance and employment, although it does not cover life insurance, disability insurance, or long-term care insurance.

Finally, there is a small chance that test results may not be accurate, leading people to make decisions based on incorrect information. Although inaccurate results are unlikely, people with these concerns should address them during genetic counseling.

What are the implications of having a harmful BRCA1 or BRCA2 mutation for breast and ovarian cancer prognosis and treatment?

A number of studies have investigated possible clinical differences between breast and ovarian cancers that are associated with harmful BRCA1 or BRCA2 mutations and cancers that are not associated with these mutations.

There is some evidence that, over the long term, women who carry these mutations are more likely to develop a second cancer in either the same (ipsilateral) breast or the opposite (contralateral) breast than women who do not carry these mutations. Thus, some women with a harmful BRCA1 or BRCA2 mutation who develop breast cancer in one breast opt for a bilateral mastectomy, even if they would otherwise be candidates for breast-conserving surgery. In fact, because of the increased risk of a second breast cancer among BRCA1 and BRCA2 mutation carriers, some doctors recommend that women with early-onset breast cancer and those whose family history is consistent with a mutation in one of these genes have genetic testing when breast cancer is diagnosed.

Breast cancers in women with a harmful BRCA1 mutation are also more likely to be "triple-negative cancers" (i.e., the breast cancer cells do not have estrogen receptors, progesterone receptors, or large amounts of HER2/neu protein), which generally have poorer prognosis than other breast cancers.

Because the products of the BRCA1 and BRCA2 genes are involved in DNA repair, some investigators have suggested that cancer cells with a harmful mutation in either of these genes may be more sensitive to anticancer agents that act by damaging DNA, such as cisplatin. In preclinical studies, drugs called PARP inhibitors, which block the repair of DNA damage, have been found to arrest the growth of cancer cells that have BRCA1 or BRCA2 mutations. These drugs have also shown some activity in cancer patients who carry BRCA1 or BRCA2 mutations, and researchers are continuing to develop and test these drugs.

What research is currently being done to help individuals with harmful BRCA1 or BRCA2 mutations?

Research studies are being conducted to find new and better ways of detecting, treating, and preventing cancer in people who carry mutations in BRCA1 and BRCA2. Additional studies are focused on improving genetic counseling methods and outcomes. Our knowledge in these areas is evolving rapidly.

Information about active clinical trials (research studies with people) for individuals with BRCA1 or BRCA2 mutations is available on NCIs website. The following links will retrieve lists of clinical trials open to individuals with BRCA1 or BRCA2 mutations.

NCIs Cancer Information Service (CIS) can also provide information about clinical trials and help with clinical trial searches.

Do inherited mutations in other genes increase the risk of breast and/or ovarian tumors?

Yes. Although harmful mutations in BRCA1 and BRCA2 are responsible for the disease in nearly half of families with multiple cases of breast cancer and up to 90 percent of families with both breast and ovarian cancer, mutations in a number of other genes have been associated with increased risks of breast and/or ovarian cancers (2, 31). These other genes include several that are associated with the inherited disorders Cowden syndrome, Peutz-Jeghers syndrome, Li-Fraumeni syndrome, and Fanconi anemia, which increase the risk of many cancer types.

Most mutations in these other genes are associated with smaller increases in breast cancer risk than are seen with mutations in BRCA1 and BRCA2. However, researchers recently reported that inherited mutations in the PALB2 gene are associated with a risk of breast cancer nearly as high as that associated with inherited BRCA1 and BRCA2 mutations (32). They estimated that 33 percent of women who inherit a harmful mutation in PALB2 will develop breast cancer by age 70 years. The estimated risk of breast cancer associated with a harmful PALB2 mutation is even higher for women who have a family history of breast cancer: 58 percent of those women will develop breast cancer by age 70 years.

PALB2, like BRCA1 and BRCA2, is a tumor suppressor gene. The PALB2 gene produces a protein that interacts with the proteins produced by the BRCA1 and BRCA2 genes to help repair breaks in DNA. Harmful mutations in PALB2 (also known as FANCN) are associated with increased risks of ovarian, pancreatic, and prostate cancers in addition to an increased risk of breast cancer (13, 33, 34). Mutations in PALB2, when inherited from each parent, can cause a Fanconi anemia subtype, FA-N, that is associated with childhood solid tumors (13, 33, 35).

Although genetic testing for PALB2 mutations is available, expert groups have not yet developed specific guidelines for who should be tested for, or the management of breast cancer risk in individuals with, PALB2 mutations.

Go here to see the original:
BRCA1 and BRCA2: Cancer Risk and Genetic Testing Fact ...

Read More...

23andMe – DNA Genetic Testing & Analysis

Sunday, October 2nd, 2016

23 pairs of chromosomes. One unique you.

Experience your ancestry in a new way! Get a breakdown of your global ancestry by percentages, connect with DNA relatives and more.

Get an even more comprehensive understanding of your genetics. Receive 65+ online reports on your ancestry, traits and health - and more.

You are made of cells. And the cells in your body have 23 pairs of chromosomes. Your chromosomes are made of DNA, which can tell you a lot about you. Explore your 23 pairs today.

Find out what your 23 pairs of chromosomes can tell you.

Your DNA analysis is performed in US laboratories that are certified to meet CLIA standardsthe Clinical Laboratory Improvement Amendments of 1988.

A CLIA-certified lab must meet certain quality standards, including qualifications for individuals who perform the test and other standards that ensure the accuracy and reliability of results.

We use leading technology to genotype your DNAa custom Illumina HumanOmniExpress-24 format chip.

Learn more about our process.

Provide your saliva sample from home. Mail it back to our lab in the same kit it came inthe postage is pre-paid.

We bring your genetics to you.

Learn more about how it works.

We hear from thousands of customers around the world who write in to tell us about their 23andMe experienceand the impact it has had on their life.

See stories that inspire us.

Here are just a few of the things people frequently ask about 23andMe. If you don't see your question here, get in touch with us.

23andMe was founded in 2006 to help people access, understand and benefit from the human genome.

We have more than one million genotyped customers around the world.

In 2015, 23andMe was granted authorization by the US Food and Drug Administration (FDA) to market the first direct-to-consumer genetic test.

23andMe offers two Personal Genetic Services: Health + Ancestry and Ancestry. Both services require submitting a saliva sample using our saliva collection kit that you send to the lab for analysis.

Our Health + Ancestry Service examines your genes to tell you about your ancestry, carrier status,* wellness and traits. We analyze, compile and distill the information extracted from your DNA into 65+ reports you can access online and share with family and friends. See full list of reports offered.

Our Ancestry Service helps you understand who you are, where you came from and your family story. We analyze, compile and distill your DNA information into reports on your Ancestry Composition, Haplogroups, Neanderthal Ancestry and provide a DNA Relatives tool to enable you to connect with people who share DNA with you.

If you have the Health + Ancestry Service you have access to the full 23andMe experience. If you only have the Ancestry Service, you can easily upgrade to the Health + Ancestry Service for $125 at any time which gives you access to all 65+ reports on ancestry, carrier status,* wellness and traits. To upgrade, log in to your 23andMe account and navigate to the Settings page. You will receive immediate access to your new health reports.

23andMe is the first and only genetic service available directly to you that includes reports that meet FDA standards for clinical and scientific validity.

Our rigorous quality standards:

You choose how your genetic information is used and shared with others. We tell you how those choices are implemented and how we collect, use and disclose your information.

*Our tests can be used to determine carrier status in adults from saliva collected using an FDA-cleared collection device (OrageneDX model OGD-500.001), but cannot determine if you have two copies of the genetic variant. The tests are not intended to diagnose a disease, or tell you anything about your risk for developing a disease in the future. On their own, carrier status tests are not intended to tell you anything about the health of your fetus, or your newborn child's risk of developing a particular disease later in life.

The Cystic Fibrosis carrier status test is indicated for the detection of 28 variants in the CFTR gene and is most relevant for people of Ashkenazi Jewish, European, and Hispanic/Latino descent. The Sickle Cell Anemia carrier status test is indicated for the detection of the HbS variant in the HBB gene and is most relevant for people of African descent. The carrier status tests related to hereditary hearing loss consist of two tests one indicated for the detection of two variants in the GJB2 gene which is most relevant for people of Ashkenazi Jewish and European descent, and one indicated for the detection of six variants in the SLC26A4 gene.

Get a breakdown of your global ancestry, connect with DNA relatives and more.

Receive 65+ online reports on your ancestry, traits and health - and more.

Your DNA can tell you about your family history. Reports include: Ancestry Composition, Haplogroups, Neanderthal Ancestry

Find and connect with relatives in the 23andMe database who share DNA with you.

If you are starting a family, find out if you are a carrier for an inherited condition. Example reports include: Cystic Fibrosis, Sickle Cell Anemia, Hereditary Hearing Loss

Learn how your genes play a role in your well-being and lifestyle choices. Example reports include: Deep Sleep, Lactose Intolerance, Saturated Fat and Weight

Explore what makes you unique, from food preferences to physical features. Example reports include: Male Bald Spot, Sweet vs. Salty, Unibrow

23andMe is the first and only genetic service available directly to you that includes reports that meet FDA standards for clinical and scientific validity.

23andMe was founded in 2006 to help people access, understand and benefit from the human genome.

We have more than one million genotyped customers around the world. Read more.

See what customers are saying about us. Read more.

You can make a difference by participating in research online, from anywhere. Read more.

We are here to help with your questions. No question is too big or small. Read more.

See the list of important policies below. Click to read more.

Your DNA can tell you about your family history. See sample report. See sample report.

Learn how your genes play a role in your well-being and lifestyle choices. See sample report. See sample report.

Explore what makes you unique, from food preference to physical features. See sample report. See sample report.

If you are starting a family, find out if you are a carrier for an inherited condition. See sample report. See sample report.

1 variant in the SACS gene; relevant for French Canadian descent

1 variant in the SLC12A6 gene; relevant for French Canadian descent

3 variants in the PKHD1 gene

10 variants in the HBB gene; relevant for Cypriot, Greek, Italian, Sardinian descent

1 variant in the BLM gene; relevant for Ashkenazi Jewish descent

2 variants in the PMM2 gene; relevant for Danish descent

28 variants in the CFTR gene; relevant for European, Hispanic/Latino, Ashkenazi Jewish descent

2 variants in the HSD17B4 gene

1 variant in the DLD gene; relevant for Ashkenazi Jewish descent

1 variant in the IKBKAP gene; relevant for Ashkenazi Jewish descent

3 variants in the FANCC gene; relevant for Ashkenazi Jewish descent

1 variant in the BCS1L gene; relevant for Finnish descent

1 variant in the G6PC gene; relevant for Ashkenazi Jewish descent

2 variants in the SLC37A4 gene

3 variants in the ALDOB gene; relevant for European descent

3 variants in the LAMB3 gene

1 variant in the LRPPRC gene; relevant for French Canadian descent

1 variant in the SGCA gene; relevant for Finnish descent

1 variant in the SGCB gene; relevant for Southern Indiana Amish descent

1 variant in the FKRP gene; relevant for European descent

3 variants in the ACADM gene; relevant for Northern European descent

2 variants in the BCKDHB gene; relevant for Ashkenazi Jewish descent

1 variant in the CLN5 gene; relevant for Finnish descent

3 variants in the PPT1 gene; relevant for Finnish descent

3 variants in the SMPD1 gene; relevant for Ashkenazi Jewish descent

1 variant in the NBN gene; relevant for Eastern European descent

2 variants in the GJB2 gene; relevant for Ashkenazi Jewish, European descent

6 variants in the SLC26A4 gene

23 variants in the PAH gene; relevant for Northern European descent

1 variant in the GRHPR gene; relevant for European descent

1 variant in the PEX7 gene

1 variant in the SLC17A5 gene; relevant for Finnish, Swedish descent

1 variant in the HBB gene; relevant for African descent

1 variant in the ALDH3A2 gene; relevant for Swedish descent

4 variants in the HEXA gene; relevant for Ashkenazi Jewish, Cajun descent

4 variants in the FAH gene; relevant for French Canadian, Finnish descent

1 variant in the PCDH15 gene; relevant for Ashkenazi Jewish descent

1 variant in the CLRN1 gene; relevant for Ashkenazi Jewish descent

1 variant in the PEX1 gene

*Our tests can be used to determine carrier status in adults from saliva collected using an FDA-cleared collection device (OrageneDX model OGD-500.001), but cannot determine if you have two copies of the genetic variant. The tests are not intended to diagnose a disease, or tell you anything about your risk for developing a disease in the future. On their own, carrier status tests are not intended to tell you anything about the health of your fetus, or your newborn child's risk of developing a particular disease later in life.

Our product is in English only, and due to the applicable regulations it is only available for customers with shipping addresses in the following countries.

If your country is not listed, visit the International site.

The rest is here:
23andMe - DNA Genetic Testing & Analysis

Read More...

preimplantation genetic diagnosis – Penn Medicine

Thursday, August 4th, 2016

All Clinical Services

Preimplantation genetic diagnosis (PGD) is a screening test used to determine if genetic or chromosomal disorders are present in embryos produced through in vitro fertilization (IVF). Preimplantation genetic diagnosis screens embryos before they are transferred to the uterus so couples can make informed decisions about their next steps in the IVF process. Embryos unaffected by the genetic or chromosomal disorder can be selected for transfer to the uterus.

For couples undergoing IVF, preimplantation genetic diagnosis may be recommended when:

Thousands of clinical preimplantation genetic diagnosis cycles have been performed worldwide, resulting in the birth of hundreds of healthy babies.

Preimplantation genetic diagnosis can be used to determine if embryos produced through in vitro fertilization carry a gene mutation associated with a specific genetic disorder, such as cystic fibrosis or muscular dystrophy.

The benefit of preimplantation genetic diagnosis is that the diagnosis can be made before the embryos are transferred to the uterus and a pregnancy is established. Embryos unaffected by the genetic disorder can be selected for transfer to the uterus, therefore greatly reducing the risk that a couple will pass a genetic disorder onto their child.

Couples who are at high risk of having a child with a severe genetic disorder may choose preimplantation genetic diagnosis for many reasons, including:

Preimplantation genetic diagnosis is also offered to couples when one partner has a chromosomal abnormality, such as an unbalanced translocation or anerplocity. If the abnormality is present in the embryo, the condition could ultimately prevent embryo implantation, lead to pregnancy loss, or result in the birth of a child with congenital malformations (physical problems) or mental retardation.

The benefit of preimplantation genetic diagnosis is that the diagnosis can be made before the embryos are transferred to the uterus and a pregnancy is established. Embryos unaffected by the chromosomal abnormality can be selected for transfer to the uterus, therefore greatly reducing the risk that the pregnancy will be adversely affected by the chromosomal abnormality.

Couples who are at high risk of having a child with a chromosomal disorder may choose preimplantation genetic diagnosis for many reasons, including:

Genetic counseling is an important step to determine if preimplantation genetic diagnosis is an appropriate option for a patient. Penn Fertility Care providers work closely with the genetic counselors in Penns Division of Reproductive Genetics. For couples undergoing IVF who are concerned that their child may inherit a genetic disorder or chromosomal abnormality, genetic counselors are available to discuss options and can advise patients on how raising a handicapped child may affect a family.

Learn more about Genetic Counseling services at Penn

Preimplantation genetic diagnosis is available for couples undergoing IVF. The steps of the IVF process include:

Embryo biopsy may be performed after 3 days of culture in the laboratory. The embryos are typically 8-cell embryos on Day-3 and the process involves the removal of one to two cells.

After the biopsy and following receipt of the results from the genetic/chromosomal testing, embryo(s) of the best quality that are not affected by the genetic disorder or chromosomal abnormality) are selected for transfer to the uterus. For day 3 embryo biopsies, the embryo is usually transferred "fresh" following two additional days of culture in the laboratory (Day-5 embryo transfer).

In some cases, the biopsy will be done on either Day-5 or -6 (trophectoderm biopsy). At this stage, the embryo consists of many cells and is called a blastocyst. Cells are removed from the outer layer of cells called the trophectoderm.

Following the biopsy of a good quality blastocyst, the blastocyst is then frozen. When the patient receives the results from the genetic testing, the non-affected or chromosomally normal blastocyst(s) are thawed and transferred in a subsequent frozen embryo cycle.

Embryos are analyzed by one of the techniques described below:

Polymerase Chain Reaction (PCR) is performed on the biopsied cell(s) to determine the presence of a single gene. This is done when a couple has a significantly increased risk of conceiving a child with a severe genetic disorder. When PCR is to be performed, the cell(s) obtained at biopsy is loaded into a tiny tube of medium and sent to for analysis. The specific area of DNA of interest is amplified by making thousands of copies of the DNA through repeated cycles of DNA strand separation and replication. The sample can be analyzed for the presence of a specific sequence of DNA or gene and also for linkage markers near the gene. The biopsied cell(s) are destroyed during this process. Therefore, they cannot be used for another purpose or returned to the embryo.

The genetic material (DNA) within the biopsied cell(s) is amplified using a technique called the polymerase chain reaction (PCR). This amplification produces enough DNA to use a second technique, known as array comparative genomic hybridization (aCGH). Array CGH assesses the amount of DNA derived from each chromosome, revealing whether or not there are both a normal amount and correct number of chromosomes. The biopsied cell(s) are destroyed during this process. Therefore, they cannot be used for another purpose or returned to the embryo. aCGH can be used to screen for numeric abnormalities in all chromosomes and/or known rearrangements of chromosomes (translocations). Array CGH does not detect all types of chromosome aberrations or genetic mutations and cannot distinguish between no translocation present and balanced translocation present.

The results of preimplantation genetic diagnosis are reported to the couple no later than the morning of their scheduled day for embryo transfer. Typically this is five days after oocyte retrieval and in vitro fertilization are performed. Of the embryo(s) that are not affected by the genetic disorder or chromosomal abnormality, the best quality embryo(s) are selected for transfer to the uterus. If additional unaffected and good-quality embryos are available, they may be cryopreserved for a future embryo transfer.

No, preimplantation genetic diagnosis does not replace prenatal testing, such as chorionic villus sampling or amniocentesis. Preimplantation genetic diagnosis provides diagnostic information based on the analysis of asinglecell. Therefore, prenatal testing is still recommended and currently remains the standard of care.

Learn more about prenatal testing services at Penn

For more information about preimplantation genetic diagnosis or to schedule an appointment with a Penn Fertility Care specialist, call 800-789-PENN (7366).

Need an appointment? Request one online 24 hours/day, 7 days/week or call 800-789-PENN (7366) to speak to a referral counselor.

Link:
preimplantation genetic diagnosis - Penn Medicine

Read More...

Genomic Medicine – Genome.gov

Thursday, August 4th, 2016

Genomic Medicine

NHGRI defines genomic medicine as "an emerging medical discipline that involves using genomic information about an individual as part of their clinical care (e.g., for diagnostic or therapeutic decision-making) and the health outcomes and policy implications of that clinical use." Already, genomic medicine is making an impact in the fields of oncology, pharmacology, rare and undiagnosed diseases, and infectious disease.

The nation's investment in the Human Genome Project (HGP) was grounded in the expectation that knowledge generated as a result of that extraordinary research effort would be used to advance our understanding of biology and disease and to improve health. In the years since the HGP's completion there has been much excitement about the potential for so-called 'personalized medicine' to reach the clinic. More recently, a report from the National Academy of Sciences [dels.nas.edu] has called for the adoption of 'precision medicine,' where genomics, epigenomics, environmental exposure, and other data would be used to more accurately guide individual diagnosis [nimh.nih.gov]. Genomic medicine, as defined above, can be considered a subset of precision medicine.

The translation of new discoveries to use in patient care takes many years. Based on discoveries over the past five to ten years, genomic medicine is beginning to fuel new approaches in certain medical specialties. Oncology, in particular, is at the leading edge of incorporating genomics, as diagnostics for genetic and genomic markers are increasingly included in cancer screening, and to guide tailored treatment strategies.

Top of page

It has often been estimated that it takes, on average, 17 years to translate a novel research finding into routine clinical practice. This time lag is due to a combination of factors, including the need to validate research findings, the fact that clinical trials are complex and take time to conduct and then analyze, and because disseminating information and educating healthcare workers about a new advance is not an overnight process.

Once sufficient evidence has been generated to demonstrate a benefit to patients, or "clinical utility," professional societies and clinical standards groups will use that evidence to determine whether to incorporate the new test into clinical practice guidelines. This determination will also factor in any potential ethical and legal issues, as well economic factors such as cost-benefit ratios.

The NHGRI Genomic Medicine Working Group (GMWG) has been gathering expert stakeholders in a series of Genomic Medicine meetings to discuss issues surrounding the adoption of genomic medicine. Particularly, the GMWG draws expertise from researchers at the cutting edge of this new medical specialty, with the aim of better informing future translational research at NHGRI. Additionally the working group provides guidance to the National Advisory Council on Human Genome Research (NACHGR) and NHGRI in other areas of genomic medicine implementation, such as outlining infrastructural needs for adoption of genomic medicine, identifying related efforts for future collaborations, and reviewing progress overall in genomic medicine implementation.

Top of page

For more examples of genomic medicine advances, please see Notable Accomplishments in Genomic Medicine

Top of page

At NHGRI, the Division of Genomic Medicine administers research programs with a clinical focus. A number of research programs currently underway are generating the evidence base, and designing and testing the implementation of genome sequencing as part of an individual's clinical care:

Within NHGRI's Division of Policy, Communications, and Education, the Policy and Program Analysis Branch (PPAB), and the Genomic Healthcare Branch (GHB) are both involved in helping pave the way for the widespread adoption of genomic medicine.

Top of page

Last Updated: March 31, 2015

Original post:
Genomic Medicine - Genome.gov

Read More...

USC Institute for Genetic Medicine

Thursday, August 4th, 2016

USC Institute for Genetic Medicine

The USC Institute for Genetic Medicine (IGM) is a place where researchers work closely together in a shared space to approach problems of the highest impact to human health. The IGM houses the offices and laboratories of a group of leading faculty members from the University of Southern California (USC). IGM scientists, engineers and clinicians take a multidisciplinary approach using molecular, genetic, omics, and computational technologies in human and genetic models.

The IGM is administered as a department reporting to the Dean of the Keck School of Medicine of USC. It is part of the USC Health Sciences Campus, located on the east side of Los Angeles.

The IGM comprises 21 faculty members. The department is an important research training ground for USC, and typically supports about 50 trainees at any one time. Including scientific and administrative staff, the total number of people working within the IGM hovers around 110 individuals.

IGM research is funded by federal, state, and industry-sponsored grants, as well as gifts from donors and benefactors. The major sponsor is the National Institutes of Health (NIH). The Center for Applied Molecular Medicine (CAMM) within the IGM holds a Physical Sciences in Oncology Center (PS-OC) grant from the NIHs National Cancer Institute.

Continued here:
USC Institute for Genetic Medicine

Read More...

Department of Genetic Medicine – Weill Cornell Medical College

Thursday, August 4th, 2016

Our translational research program includes many projects in the fields of genetic therapies and personalized medicine. The field of genetic therapies comprises gene and stem cell therapies and our laboratory has extensive expertise in both areas. Our group was the first to use a recombinant virus as a vehicle for in vivo gene therapy and we have carried out human trials of gene therapy for cystic fibrosis, cardiac ischemia, cancer and central nervous system disorders. Among the current projects are gene transfer strategies for cancer, inherited CNS disorders, 1-antitrypsin deficiency, anti-bioterrorism applications and development of vaccines. We also operate the clinical vector production laboratory of the Belfer Gene Therapy Core Facility, which has produced adenovirus and adeno-associated virus vectors that have been used in numerous human studies. Current projects in the field of stem cell therapy include characterization of the roles of cancer stem cells in lung cancer and the role of airway epithelium stem cells in chronic obstructive pulmonary disease.

Personalized medicine is the use of information and data from an individual's genotype, or level of gene expression to stratify complex diseases, select a medication or dose of a medication, provide a therapy, or initiate a preventative measure that is specifically suited to that patient. In addition to genetic information, other factors, including imaging, laboratory, and clinical information about the disease process or the patient are integrated into the process of developing personalized medicine. Our group utilizes microarray technologies for genome-wide characterization of gene expression, single nucleotide polymorphism and copy number variation profiles on clinical samples as the basis for projects aimed at indentifying candidate genes associated with complex disease such as chronic obstructive pulmonary disease.

The overall research program of the group includes close collaborations with other laboratories at Weill Cornell and elsewhere, including Malcolm Moore's group at Memorial Sloan Kettering Cancer Center for stem cell projects. Of particular note are our collaborations on personalized medicine projects with colleagues at Weill Cornell Medical College-Qatar and Hamad Medical Corporation in Doha, Qatar and collaborations on Bioinformatics and Biostatistical Genetics with several laboratories at Cornell-Ithaca, including Andy Clark and Jason Mezey.

The rest is here:
Department of Genetic Medicine - Weill Cornell Medical College

Read More...

Home: Feinberg School of Medicine: Northwestern University

Thursday, August 4th, 2016

How to manage stress when the world is filled with worry

A little stress can actually be a good thing, motivating us to work hard and get ahead, experts say. But constant stress and worry over the long haul can damage our bodies. "The stress response was made for short-term acute stress, like needing to run away from a bear or a saber tooth tiger," said David Victorson, an associate professor of medical social sciences at the Northwestern University Feinberg School of Medicine and a health psychologist at Northwestern Medicine. "It's been a part of the human process since the beginning. But stressors today can be much more chronic and we're ill equipped to deal with that.

Garfield, an associate professor of pediatrics at the Northwestern University Feinberg School of Medicine, said research on fatherhood is a fairly recent phenomenon, so it's difficult to compare dads of today with fathers in the '60s, for example. But he said change is in the air, as evidenced not only by formal studies, but by cultural phenomena such as the rise of 'dad-vertising,' in which fathers are portrayed as capable, hands-on parents, rather than workaholics or bumbling oafs.

Researchers in Illinois have unveiled the third gene linked with Parkinsons, a discovery that comes following the death of legendary boxer Muhammad Ali, who suffered from the neurodegenerative disease for three decades. Scientists findings, published Monday in Nature Genetics, suggest the genetic mutation TMEM230 was present among Parkinsons patients in North America and Asia, and had similar protein trafficking characteristics as the other two genetic mutations linked with Parkinsons, according to a Northwestern University press release. They found TMEM230 produced a protein involved in the packaging of dopamine in neurons, which is significant because Parkinsons is marked by the breakdown of dopamine-producing neurons.

Not getting a good night's sleep can result in a number of problems including poor concentration, weight gain, and a greater likelihood of accidents. For shift workers and individuals who experience chronic sleep deprivation, new research suggests insufficient sleep could also increase the risk of heart disease. In humans, as in all mammals, almost all physiological and behavioral processes, in particular the sleep-wake cycle, follow a circadian rhythm that is regulated by an internal clock located in the brain," said Daniela Grimaldi, M.D., Ph.D., lead author and a research assistant professor at Northwestern University, said in a press release. "When our sleep-wake and feeding cycles are not in tune with the rhythms dictated by our internal clock, circadian misalignment occurs."

Low-income families with children with allergies spend more than twice as much on visits to emergency rooms and hospitals than mid- to high-income families, recent research from Northwestern University found. And about 40 percent of those children surveyed also reported experiencing life-threatening reactions to food, such as trouble breathing and a drop in blood pressure. "The fact that they were able to open up a food pantry for kids who can't afford the special foods for food allergies incredible," said Dr. Ruchi Gupta, an associate professor of pediatrics who led the Northwestern study, which was published in April.

We need a lot more nonsteroidal options, and [crisaborole] looks like it may be an important addition to our armamentarium, says Jonathan Silverberg, a dermatologist at Northwestern University Feinberg School of Medicine. We have limited options for the thing we can safely give patients without worries about their long term use.

Go here to see the original:
Home: Feinberg School of Medicine: Northwestern University

Read More...

National Human Genome Research Institute (NHGRI) – Homepage

Thursday, August 4th, 2016

New NIH studies seek adults and families affected by sickle cell disease/trait People with sickle cell disease (SCD) can experience excruciating pain, kidney problems, a higher risk of stroke and, in rare cases, chronic leg ulcers. Little is known about why the severity of these symptoms varies throughout a lifetime or why these symptoms differ from person to person. NHGRI researchers are seeking help from people affected by SCD to find the factors - environmental, social and genetic - that impact the severity of the symptoms. Read more Investigational Device Exemptions (IDE) and Genomics Workshop On Friday, June 10, 2016, the National Human Genome Research Institute (NHGRI) hosted the Investigational Device Exemptions (IDE) and Genomics Workshop. The workshop brought together perspectives from investigators, institutional review boards (IRB), the FDA and NHGRI on how to determine if a study requires an IDE, and how to fulfill IDE requirements if the FDA should require an IDE for research involving the use of genomic technologies, including next-generation sequencing (NGS). View agenda The Genomics Landscape Clinical Sequencing: Beyond Exploration The Genomics Landscape for June features exciting developments with NHGRI's Clinical Sequencing Exploratory Research Program. We also highlight the new director of the National Library of Medicine, recently funded studies on the ethical, legal and social implications of genomic information, the final seminar commemorating the 25th anniversary of the launch of the Human Genome Project and available online videos for genome analysis lectures. Read more Bacterial toxins make the body go boom By outward appearances, plants and animals don't seem to have much in common. When it comes to their immune systems, however, they might be more similar than their exteriors suggest. Researchers at the National Human Genome Research Institute have discovered an immune mechanism in humans, known as a "guard" mechanism, which was once thought to exist only in plants. They've published their results in the June 6 online journal Nature Immunology. Read more Perspective: Precision medicine may move us beyond the use of race in prescribing drugs Health care providers have long struggled with considering race when prescribing and dosing medications. In a May 26 New England Journal of Medicine perspective, Vence L. Bonham, J.D., an investigator with NHGRI's Social and Behavioral Research Branch, and his colleagues, are asking if the precision medicine approach will reduce or eliminate the role that race plays in prescribing drugs and in health care overall. Read more

See original here:
National Human Genome Research Institute (NHGRI) - Homepage

Read More...

What is Genomic Medicine? – Genome.gov

Thursday, August 4th, 2016

What is Genomic Medicine?

NHGRI defines genomic medicine as "an emerging medical discipline that involves using genomic information about an individual as part of their clinical care (e.g., for diagnostic or therapeutic decision-making) and the health outcomes and policy implications of that clinical use." Already, genomic medicine is making an impact in the fields of oncology, pharmacology, rare and undiagnosed diseases, and infectious disease.

The nation's investment in the Human Genome Project (HGP) was grounded in the expectation that knowledge generated as a result of that extraordinary research effort would be used to advance our understanding of biology and disease and to improve health. In the years since the HGP's completion there has been much excitement about the potential for so-called 'personalized medicine' to reach the clinic. More recently, a report from the National Academy of Sciences [dels.nas.edu] has called for the adoption of 'precision medicine,' where genomics, epigenomics, environmental exposure, and other data would be used to more accurately guide individual diagnosis [nimh.nih.gov]. Genomic medicine, as defined above, can be considered a subset of precision medicine.

The translation of new discoveries to use in patient care takes many years. Based on discoveries over the past five to ten years, genomic medicine is beginning to fuel new approaches in certain medical specialties. Oncology, in particular, is at the leading edge of incorporating genomics, as diagnostics for genetic and genomic markers are increasingly included in cancer screening, and to guide tailored treatment strategies.

Top of page

It has often been estimated that it takes, on average, 17 years to translate a novel research finding into routine clinical practice. This time lag is due to a combination of factors, including the need to validate research findings, the fact that clinical trials are complex and take time to conduct and then analyze, and because disseminating information and educating healthcare workers about a new advance is not an overnight process.

Once sufficient evidence has been generated to demonstrate a benefit to patients, or "clinical utility," professional societies and clinical standards groups will use that evidence to determine whether to incorporate the new test into clinical practice guidelines. This determination will also factor in any potential ethical and legal issues, as well economic factors such as cost-benefit ratios.

The NHGRI Genomic Medicine Working Group (GMWG) has been gathering expert stakeholders in a series of Genomic Medicine meetings to discuss issues surrounding the adoption of genomic medicine. Particularly, the GMWG draws expertise from researchers at the cutting edge of this new medical specialty, with the aim of better informing future translational research at NHGRI. Additionally the working group provides guidance to the National Advisory Council on Human Genome Research (NACHGR) and NHGRI in other areas of genomic medicine implementation, such as outlining infrastructural needs for adoption of genomic medicine, identifying related efforts for future collaborations, and reviewing progress overall in genomic medicine implementation.

Top of page

For more examples of genomic medicine advances, please see Notable Accomplishments in Genomic Medicine

Top of page

At NHGRI, the Division of Genomic Medicine administers research programs with a clinical focus. A number of research programs currently underway are generating the evidence base, and designing and testing the implementation of genome sequencing as part of an individual's clinical care:

Within NHGRI's Division of Policy, Communications, and Education, the Policy and Program Analysis Branch (PPAB), and the Genomic Healthcare Branch (GHB) are both involved in helping pave the way for the widespread adoption of genomic medicine.

Top of page

Last Updated: March 31, 2015

See the original post here:
What is Genomic Medicine? - Genome.gov

Read More...

Department of Genetic Medicine (Research) | – | Weill …

Thursday, November 5th, 2015

The Department of Genetic Medicine at Weill Cornell Medicine is a highly specialized form of personalized medicine that involves the introduction of genetic material into a patients cells to fight or prevent disease. This experimental approach requires the use of information and data from an individual's genotype or specific DNA signature, to challenge a disease, select a medication or its dosage, provide a specific therapy, or initiate preventative measures specifically suited to the patient. While this technology is still in its infancy, gene therapy has been used with some success and offers the promise of regenerative cures.

As none of New York's premier healthcare networks, Weill Cornell Medicine's genetic research program includes close collaborations with fellow laboratories such as Memorial Sloan Kettering Cancer Center for stem cell projects, Weill Cornell Medical College in Qatar and Hamad Medical Corporation in Doha, Qatar and Bioinformatics and Biostatistical Genetics at Cornell-Ithaca.

Department of Genetic Medicine Services

Our translational research program includes many projects in the fields of genetic therapies and personalized medicine, and we arestudying gene therapy for a number of diseases, such as combined immuno-deficiencies, hemophilia, Parkinson's, cancer and even HIV using a number of different approaches.

Patients interested in gene therapy are invited to participate in our full range of services, including:

-diagnostic testing

-imaging

-laboratory analysis

-clinical informatics

-managed therapies

In addition, we offer genetic testing to provide options for individuals and families seeking per-emptive strategies for addressing the uncertainties surrounding inherited diseases.The Department of Genetic Medicine at Weill Cornell is a pioneer in the advancement of genetics for patients and their families. These are the strengths we draw upon as we collaborate with our integrated network of partners, including the #1 hospital in New York, New York Presbyterian, to make breakthroughs a reality for our patients.

For more information or to schedule an appointment, call us toll-free at 1-855-WCM-WCMU.

Go here to see the original:
Department of Genetic Medicine (Research) | - | Weill ...

Read More...

Graduate Program in Genetic Counseling : Center for …

Thursday, October 8th, 2015

Northwestern University provides a strong foundation in core genetic counseling skills and identifies each students strengths in order to ignite the passion and lifelong commitment to learning that is critical to professional development. Graduates not only feel extremely capable in multiple clinical settings and specialties, but also recognize how valuable their training has been in preparing them for expanded genetic counseling careers.

Since the inception of the Northwestern University Graduate Program in Genetic Counseling in 1990, the leaders of the program have strived to look to the future of the genetic counseling profession to help guide the overall administration and curriculum. The field of genetics has evolved rapidly over time, and graduate programs need to be aware of the changes that will continue to shape and influence the profession. Northwestern has continued to successfully evolve to meet these changing needs. There are several strengths that allow Northwestern to maintain this cutting edge:

This unique combination, along with the personalized attention a student receives during their training, creates an exciting learning environment and is one of the major strengths of the Northwestern program. We believe our students deserve a strong science, research and psychosocial curriculum.

In addition, Northwestern is proud to offer one of the only dual degree programs available in Genetic Counseling and Medical Humanities and Bioethics.

The combination of the programs nationally recognized faculty, the diversity of clinical and patient experiences, and the cultural excitement of its location in Chicago makes this program unique, exciting and visionary!

Learn more about the program via the links below.

View post:
Graduate Program in Genetic Counseling : Center for ...

Read More...

Genetic Counseling Program – University of South Carolina …

Friday, September 25th, 2015

The two year curriculum includes course work, clinical rotations, and a research-based thesis. Students are afforded a wide range of clinical opportunities in prenatal, pediatric and adult settings as well as specialty clinics through our clinical rotation network. International rotations are encouraged.

In 1991 and 1998, the Program received rare Commendation for Excellence citations from the South Carolina Commission of Higher Education. The Program was awarded American Board of Genetic Counseling accreditation in 2000 and reaccreditation in 2006. Most recently, the Accreditation Council for Genetic Counseling re-accredited the Program for the maximum eight year period, 2014-2022.

We invite you to explore the University of South Carolina Genetic Counseling Program through this site. Please also visit the National Society of Genetic Counselors, the American Board of Genetic Counseling websites to learn more about the profession. Check out the latest U.S. Department of Labor, Occupational Outlook Handbook, 2014-15 Edition projections for genetic counselors. The future is bright for genetic counselors!

Go here to read the rest:
Genetic Counseling Program - University of South Carolina ...

Read More...

Genome Medicine

Monday, September 7th, 2015

Medicine in the post-genomic era

Genome Medicine publishes peer-reviewed research articles, new methods, software tools, reviews and comment articles in all areas of medicine studied from a post-genomic perspective. Areas covered include, but are not limited to, disease genomics (including genome-wide association studies and sequencing-based studies), disease epigenomics, pathogen and microbiome genomics, immunogenomics, translational genomics, pharmacogenomics and personalized medicine, proteomics and metabolomics in medicine, systems medicine, and ethical, legal and social issues.

There has been an error retrieving the data. Please try again.

DNA-PK inhibition boosts Cas9-mediated HDR

Transient pharmacological inhibition of DNA-PKcs can stimulate homology-directed repair following Cas9-mediated induction of a double strand break, and is expected to reduce the downstream workload.

Genomics of epilepsy

Candace Myers and Heather Mefford review how advances in genomic technologies have aided variant discovery, leading to a rapid increase in our understanding of epilepsy genetics.

CpG sites associated with atopy

Thirteen novel epigenetic loci associated with atopy and high IgE were found that could serve 55 as candidate loci; of these, four were within genes with known roles in the immune response.

Longitudinal 'omic profiles

A pilot study quantifying gene expression and methylation profile consistency over a year shows high longitudinal consistency, with individually extreme transcript abundance in a small number of genes which may be useful for explaining medical conditions or guiding personalized health decisions.

Ovarian cancer landscape

Exome sequencing of mucinous ovarian carcinoma tumors reveals multiple mutational targets, suggesting tumors arise through many routes, and shows this group of tumors is distinct from other subtypes.

NGS-guided cancer therapy

Jeffrey Gagan and Eliezer Van Allen review how next-generation sequencing can be incorporated into standard oncology clinical practice and provide guidance on the potential and limitations of sequencing.

ClinLabGeneticist

A platform for managing clinical exome sequencing data that includes data entry, distribution of work assignments, variant evaluation and review, selection of variants for validation, report generation.

Semantic workflow for clinical omics

A clinical omics analysis pipeline using the Workflow Instance Generation and Specialization (WINGS) semantic workflow platform demonstrates transparency, reproducibility and analytical validity.

Stephen McMahon and colleagues review treatments for pain relief, which are often inadequate, and discuss how understanding of the genomic and epigenomic mechanisms might lead to improved drugs.

View more review articles

Exploiting single-molecule transcript sequencing for eukaryotic gene prediction

Minoche AE, Dohm JC, Schneider J, Holtgrwe D, Viehver P, Montfort M, Rosleff Srensen T, Weisshaar B et al.

Genome Biology 2015, 16:184

Analysis methods for studying the 3D architecture of the genome

Ay F and Noble WS

Genome Biology 2015, 16:183

Graded gene expression changes determine phenotype severity in mouse models of CRX-associated retinopathies

Ruzycki PA, Tran NM, Kefalov VJ, Kolesnikov AV and Chen S

Genome Biology 2015, 16:171

Predicting the spatial organization of chromosomes using epigenetic data

Mourad R and Cuvier O

Genome Biology 2015, 16:182

See the original post here:
Genome Medicine

Read More...

NIH Clinical Center: Graduate Medical Education (GME …

Friday, August 21st, 2015

Graduate Medical Education (GME): Medical Genetics

Maximilian Muenke, MD

Eligibility CriteriaCandidates with the MD degree must have completed an accredited U.S. residency training program and have a valid U.S. license. Previous training is usually in, but not limited to, Pediatrics, Internal Medicine or Obstetrics and Gynecology.

OverviewThe NIH has joined forces with training programs at the Children's National Medical Center, George Washington University School of Medicine and Washington Hospital Center. The combined training program in Medical Genetics is called the Metropolitan Washington, DC Medical Genetics Program. This is a program of three years duration for MDs seeking broad exposure to both clinical and research experience in human genetics.

The NIH sponsor of the program is National Human Genome Research Institute (NHGRI). Other participating institutes include the National Cancer Institute (NCI), the National Eye Institute (NEI), the National Institute of Arthritis and Musculoskeletal and Skin Diseases (NIAMS), the National Institute of Child Health and Human Development (NICHD), the National Institute on Deafness and Other Communication Disorders (NIDCD), the National Institute of Diabetes and Digestive and Kidney Diseases (NIDDK), and the National Institute of Mental Health (NIMH). Metropolitan area participants include Children's National Medical Center (George Washington University), Walter Reed Army Medical Center, and the Department of Pediatrics, and the Department of Obstetrics and Gynecology at Washington Hospital Center. The individual disciplines in the program include clinical genetics, biochemical genetics, clinical cytogenetics, and clinical molecular genetics.

The primary goal of the training program is to provide highly motivated physicians with broad exposure to both clinical and research experiences in medical genetics. We train candidates to become effective, independent medical geneticists, prepared to deliver a high standard of clinical genetics services, and to perform state-of-the-art research in the area of genetic disease.

Structure of the Clinical Training Program

RotationsThis three year program involves eighteen months devoted to learning in clinical genetics followed by eighteen months of clinical or laboratory research.

Year 1Six months will be spent on rotation at the NIH. Service will include time spent on different outpatient genetics clinics, including Cancer Genetics and Endocrine Disorders and Genetic Ophthalmology; on the inpatient metabolic disease and endocrinology ward; on inpatient wards for individuals involved in gene therapy trials; and on the NIH Genetics Consultation Service.

Three months will be spent at Children's National Medical Center and will be concentrated on pediatric genetics. Fellows will participate in outpatient clinics, satellite and outreach clinics. They will perform consults on inpatients and patients with metabolic disorders and on the neonatal service. Fellows will be expected to participate in the relevant diagnostic laboratory studies on patients for whom they have provided clinical care.

One month will be spent at Walter Reed Army Medical Center and will concentrate on adult and pediatric clinical genetics. One month will be spent at Washington Hospital Center on rotations in prenatal genetics and genetic counseling.

Year 2 Fellows will spend one month each in clinical cytogenetics, biochemical genetics, and molecular diagnostic laboratories. The remaining three months will be devoted to elective clinical rotations on any of the rotations previously mentioned. The second six months will be spent on laboratory or clinical research. The fellow will spend at least a half-day per week in clinic at any one of the three participating institutions.

Year 3This year will be devoted to research, with at least a half day per week in clinic.

NIH Genetics Clinic (Required)Fellows see patients on a variety of research protocols. The Genetics Clinic also selectively accepts referrals of patients requiring diagnostic assessment and genetic counseling. Areas of interest and expertise include: chromosomal abnormalities, congenital anomalies and malformation syndromes, biochemical defects, bone and connective tissue disorders, neurological disease, eye disorders, and familial cancers.

Inpatient Consultation Service (Required)Fellows are available twenty-four hours daily to respond to requests for genetics consultation throughout the 325-bed hospital. Written consultation procedures call for a prompt preliminary evaluation, a written response within twenty-four hours, and a subsequent presentation to a senior staff geneticist, with an addendum to the consult, as needed. The consultant service fellow presents the most interesting cases from the wards during the Post-Clinic Patient Conference on Wednesday afternoons during which Fellows present interesting clinical cases for critical review. Once a month the fellow presents relevant articles for journal club.

Metropolitan Area Genetics Clinics

Other Clinical Opportunities: Specialty Clinics at NIHThe specialty clinics of NIH treat a large number of patients with genetic diseases. We have negotiated a supervised experience for some of the fellows at various clinics; to date, fellows have participated in the Cystic Fibrosis Clinic, the Lipid Clinic, and the Endocrine Clinic.

Lectures, Courses and SeminarsThe fellowship program includes many lectures, courses and seminars. Among them are a journal club and seminars in medical genetics during which invited speakers discuss research and clinical topics of current interest. In addition, the following four courses have been specifically developed to meet the needs of the fellows:

Trainees are encouraged to pursue other opportunities for continuing education such as clinical and basic science conferences, tutorial seminars, and postgraduate courses, which are plentiful on the NIH campus.

Structure of the Research Training ProgramFellows in the Medical Genetics Program pursue state-of-the-art research related to genetic disorders. Descriptions of the diverse interests of participating faculty are provided in this catalog. The aim of this program is to provide fellows with research experiences of the highest caliber and to prepare them for careers as independent clinicians and investigators in medical genetics.

Fellows entering the program are required to select a research supervisor which may be from among those involved on the Genetics Fellowship Faculty Program. It is not required that this selection be made before coming to NIH.

In addition to being involved in research, all fellows attend and participate in weekly research seminars, journal clubs and laboratory conferences, which are required elements of each fellow's individual research experience.

Program Faculty and Research Interests

Examples of Papers Authored by Program Faculty

Program GraduatesThe following is a partial list of graduates including their current positions:

Application Information

The NIH/Metropolitan Washington Medical Genetics Residency Program is accredited by the ACGME and the American Board of Medical Genetics. Upon successful completion of the three year program, residents are eligible for board certification in Clinical Genetics. During the third residency year, residents may elect to complete either (a) the requirements for one of the ABMG laboratory subspecialties, such as Clinical Molecular Genetics, Clinical Biochemical Genetics or Clinical Cytogenetics, or (b) a second one year residency program (e.g., Medical Biochemical Genetics).

Candidates should apply through ERAS, beginning July 1 of the year prior to their anticipated start date. Candidates with the MD or MD and PhD degree must have completed a U.S. residency in a clinically related field. Previous training is usually in, but not limited to, Pediatrics, Internal Medicine or Obstetrics and Gynecology. Four new positions are available each year. Interviews are held during August and September.

Electronic Application The quickest and easiest way to find out more about this training program or to apply for consideration is to do it electronically.

The NIH is dedicated to building a diverse community in its training and employment programs.

NOTE: PDF documents require the free Adobe Reader.

View post:
NIH Clinical Center: Graduate Medical Education (GME ...

Read More...

Genetics & Medicine – National Center for Biotechnology …

Saturday, August 15th, 2015

Bookshelf

A collection of biomedical books that can be searched directly or from linked data in other NCBI databases. The collection includes biomedical textbooks, other scientific titles, genetic resources such as GeneReviews, and NCBI help manuals.

A resource to provide a public, tracked record of reported relationships between human variation and observed health status with supporting evidence. Related information intheNIH Genetic Testing Registry (GTR),MedGen,Gene,OMIM,PubMedand other sources is accessible through hyperlinks on the records.

An archive and distribution center for the description and results of studies which investigate the interaction of genotype and phenotype. These studies include genome-wide association (GWAS), medical resequencing, molecular diagnostic assays, as well as association between genotype and non-clinical traits.

An open, publicly accessible platform where the HLA community can submit, edit, view, and exchange data related to the human major histocompatibility complex. It consists of an interactive Alignment Viewer for HLA and related genes, an MHC microsatellite database, a sequence interpretation site for Sequencing Based Typing (SBT), and a Primer/Probe database.

A searchable database of genes, focusing on genomes that have been completely sequenced and that have an active research community to contribute gene-specific data. Information includes nomenclature, chromosomal localization, gene products and their attributes (e.g., protein interactions), associated markers, phenotypes, interactions, and links to citations, sequences, variation details, maps, expression reports, homologs, protein domain content, and external databases.

A collection of expert-authored, peer-reviewed disease descriptions on the NCBI Bookshelf that apply genetic testing to the diagnosis, management, and genetic counseling of patients and families with specific inherited conditions.

Summaries of information for selected genetic disorders with discussions of the underlying mutation(s) and clinical features, as well as links to related databases and organizations.

A voluntary registry of genetic tests and laboratories, with detailed information about the tests such as what is measured and analytic and clinical validity. GTR also is a nexus for information about genetic conditions and provides context-specific links to a variety of resources, including practice guidelines, published literature, and genetic data/information. The initial scope of GTR includes single gene tests for Mendelian disorders, as well as arrays, panels and pharmacogenetic tests.

A database of known interactions of HIV-1 proteins with proteins from human hosts. It provides annotated bibliographies of published reports of protein interactions, with links to the corresponding PubMed records and sequence data.

A compilation of data from the NIAID Influenza Genome Sequencing Project and GenBank. It provides tools for flu sequence analysis, annotation and submission to GenBank. This resource also has links to other flu sequence resources, and publications and general information about flu viruses.

A portal to information about medical genetics. MedGen includes term lists from multiple sources and organizes them into concept groupings and hierarchies. Links are also provided to information related to those concepts in the NIH Genetic Testing Registry (GTR), ClinVar,Gene, OMIM, PubMed, and other sources.

A database of human genes and genetic disorders. NCBI maintains current content and continues to support its searching and integration with other NCBI databases. However, OMIM now has a new home at omim.org, and users are directed to this site for full record displays.

A database of citations and abstracts for biomedical literature from MEDLINE and additional life science journals. Links are provided when full text versions of the articles are available via PubMed Central (described below) or other websites.

A digital archive of full-text biomedical and life sciences journal literature, including clinical medicine and public health.

A collection of clinical effectiveness reviews and other resources to help consumers and clinicians use and understand clinical research results. These are drawn from the NCBI Bookshelf and PubMed, including published systematic reviews from organizations such as the Agency for Health Care Research and Quality, The Cochrane Collaboration, and others (see complete listing). Links to full text articles are provided when available.

A collection of resources specifically designed to support the research of retroviruses, including a genotyping tool that uses the BLAST algorithm to identify the genotype of a query sequence; an alignment tool for global alignment of multiple sequences; an HIV-1 automatic sequence annotation tool; and annotated maps of numerous retroviruses viewable in GenBank, FASTA, and graphic formats, with links to associated sequence records.

A summary of data for the SARS coronavirus (CoV), including links to the most recent sequence data and publications, links to other SARS related resources, and a pre-computed alignment of genome sequences from various isolates.

An extension of the Influenza Virus Resource to other organisms, providing an interface to download sequence sets of selected viruses, analysis tools, including virus-specific BLAST pages, and genome annotation pipelines.

Go here to see the original:
Genetics & Medicine - National Center for Biotechnology ...

Read More...

Medical genetics of Jews – Wikipedia, the free encyclopedia

Saturday, August 15th, 2015

The medical genetics of Jews is the study, screening, and treatment of genetic disorders more common in particular Jewish populations than in the population as a whole.[1] The genetics of Ashkenazi Jews have been particularly well-studied, resulting in the discovery of many genetic disorders associated with this ethnic group. In contrast, the medical genetics of Sephardic Jews and Mizrahi Jews are more complicated, since they are more genetically diverse and consequently no genetic disorders are more common in these groups as a whole; instead, they tend to have the genetic diseases common in their various countries of origin.[1][2] Several organizations, such as Dor Yeshorim,[3] offer screening for Ashkenazi genetic diseases, and these screening programs have had a significant impact, in particular by reducing the number of cases of TaySachs disease.[4]

Different ethnic groups tend to suffer from different rates of hereditary diseases, with some being more common, and some less common. Hereditary diseases, particularly hemophilia, were recognized early in Jewish history, even being described in the Talmud.[5] However, the scientific study of hereditary disease in Jewish populations was initially hindered by scientific racism, which believed in racial supremacism.[6][7]

However, modern studies on the genetics of particular ethnic groups have the tightly defined purpose of avoiding the birth of children with genetic diseases, or identifying people at particular risk of developing a disease in the future.[6] Consequently, the Jewish community has been very supportive of modern genetic testing programs, although this unusually high degree of cooperation has raised concerns that it might lead to the false perception that Jews are more susceptible to genetic diseases than other groups of people.[5]

However, most populations contain hundreds of alleles that could potentially cause disease and most people are heterozygotes for one or two recessive alleles that would be lethal in a homozygote.[8] Although the overall frequency of disease-causing alleles does not vary much between populations, the practice of consanguineous marriage (marriage between second cousins or closer relatives) is common in some Jewish communities, which produces a small increase in the number of children with congenital defects.[9]

According to Daphna Birenbaum Carmeli at the University of Haifa, Jewish populations have been studied more thoroughly than most other human populations because:[10]

The result is a form of ascertainment bias. This has sometimes created an impression that Jews are more susceptible to genetic disease than other populations. Carmeli writes, "Jews are over-represented in human genetic literature, particularly in mutation-related contexts."[10] Another factor that may aid genetic research in this community is that Jewish culture results in excellent medical care, which is coupled to a strong interest in the community's history and demography.[11]

This set of advantages have led to Ashkenazi Jews in particular being used in many genetic studies, not just in the study of genetic diseases. For example, a series of publications on Ashkenazi centenarians established their longevity was strongly inherited and associated with lower rates of age-related diseases.[12] This "healthy aging" phenotype may be due to higher levels of telomerase in these individuals.[13]

The most detailed genetic analysis study of Ashkenazi was published in September 2014 by Shai Carmon and his team at Columbia University. The results of the detailed study show that today's 10 million Ashkenai Jews descend from a population only 350 individuals who lived about 600-800 years ago. That population derived from both Europe and the Middle East. [14]There is evidence that the population bottleneck may have allowed deleterious alleles to become more prevalent in the population due to genetic drift.[15] As a result, this group has been particularly intensively studied, so many mutations have been identified as common in Ashkenazis.[16] Of these diseases, many also occur in other Jewish groups and in non-Jewish populations, although the specific mutation which causes the disease may vary between populations. For example, two different mutations in the glucocerebrosidase gene causes Gaucher's disease in Ashkenazis, which is their most common genetic disease, but only one of these mutations is found in non-Jewish groups.[4] A few diseases are unique to this group; for example, familial dysautonomia is almost unknown in other populations.[4]

TaySachs disease, a fatal illness of children that causes mental deterioration prior to death, was historically more prevalent among Ashkenazi Jews,[18] although high levels of the disease are also found in some Pennsylvania Dutch, southern Louisiana Cajun, and eastern Quebec French Canadian populations.[19] Since the 1970s, however, proactive genetic testing has been quite effective in eliminating TaySachs from the Ashkenazi Jewish population.[20]

Gaucher's disease, in which lipids accumulate in inappropriate locations, occurs most frequently among Ashkenazi Jews;[21] the mutation is carried by roughly one in every 15 Ashkenazi Jews, compared to one in 100 of the general American population.[22] Gaucher's disease can cause brain damage and seizures, but these effects are not usually present in the form manifested among Ashkenazi Jews; while sufferers still bruise easily, and it can still potentially rupture the spleen, it generally has only a minor impact on life expectancy.

Ashkenazi Jews are also highly affected by other lysosomal storage diseases, particularly in the form of lipid storage disorders. Compared to other ethnic groups, they more frequently act as carriers of mucolipidosis[23] and NiemannPick disease,[24] the latter of which can prove fatal.

The occurrence of several lysosomal storage disorders in the same population suggests the alleles responsible might have conferred some selective advantage in the past.[25] This would be similar to the hemoglobin allele which is responsible for sickle-cell disease, but solely in people with two copies; those with just one copy of the allele have a sickle cell trait and gain partial immunity to malaria as a result. This effect is called heterozygote advantage.[26]

Some of these disorders may have become common in this population due to selection for high levels of intelligence (see Ashkenazi intelligence).[27][28] However, other research suggests no difference is found between the frequency of this group of diseases and other genetic diseases in Ashkenazis, which is evidence against any specific selectivity towards lysosomal disorders.[29]

Familial dysautonomia (RileyDay syndrome), which causes vomiting, speech problems, an inability to cry, and false sensory perception, is almost exclusive to Ashkenazi Jews;[30] Ashkenazi Jews are almost 100 times more likely to carry the disease than anyone else.[31]

Diseases inherited in an autosomal recessive pattern often occur in endogamous populations. Among Ashkenazi Jews, a higher incidence of specific genetic disorders and hereditary diseases have been verified, including:

In contrast to the Ashkenazi population, Sephardic and Mizrahi Jews are much more divergent groups, with ancestors from Spain, Portugal, Morocco, Tunisia, Algeria, Italy, Libya, the Balkans, Iran, Iraq, India, and Yemen, with specific genetic disorders found in each regional group, or even in specific subpopulations in these regions.[1]

One of the first genetic testing programs to identify heterozygote carriers of a genetic disorder was a program aimed at eliminating TaySachs disease. This program began in 1970, and over one million people have now been screened for the mutation.[46] Identifying carriers and counseling couples on reproductive options have had a large impact on the incidence of the disease, with a decrease from 4050 per year worldwide to only four or five per year.[4] Screening programs now test for several genetic disorders in Jews, although these focus on the Ashkenazi Jews, since other Jewish groups cannot be given a single set of tests for a common set of disorders.[2] In the USA, these screening programs have been widely accepted by the Ashkenazi community, and have greatly reduced the frequency of the disorders.[47]

Prenatal testing for several genetic diseases is offered as commercial panels for Ashkenazi couples by both CIGNA and Quest Diagnostics. The CIGNA panel is available for testing for parental/preconception screening or following chorionic villus sampling or amniocentesis and tests for Bloom syndrome, Canavan disease, cystic fibrosis, familial dysautonomia, Fanconi anemia, Gaucher disease, mucolipidosis IV, Neimann-Pick disease type A, Tay-Sachs disease, and torsion dystonia. The Quest panel is for parental/preconception testing and tests for Bloom syndrome, Canavan disease, cystic fibrosis, familial dysautonomia, Fanconi anemia group C, Gaucher disease, Neimann-Pick disease types A and B and Tay-Sachs disease.

The official recommendations of the American College of Obstetricians and Gynecologists is that Ashkenazi individuals be offered screening for Tay Sachs, Canavan, cystic fibrosis, and familial dysautonomia as part of routine obstetrical care.[48]

In the orthodox community, an organization called Dor Yeshorim carries out anonymous genetic screening of couples before marriage to reduce the risk of children with genetic diseases being born.[49] The program educates young people on medical genetics and screens school-aged children for any disease genes. These results are then entered into an anonymous database, identified only by a unique ID number given to the person who was tested. If two people are considering getting married, they call the organization and tell them their ID numbers. The organization then tells them if they are genetically compatible. It is not divulged if one member is a carrier, so as to protect the carrier and his or her family from stigmatization.[49] However, this program has been criticized for exerting social pressure on people to be tested, and for screening for a broad range of recessive genes, including disorders such as Gaucher's disease.[3]

Related:

Read the original:
Medical genetics of Jews - Wikipedia, the free encyclopedia

Read More...

Genetic Disease: Get the Definition of These Disorders

Friday, July 31st, 2015

What is a genetic disease?

A genetic disease is any disease that is caused by an abnormality in an individual's genome, the person's entire genetic makeup. The abnormality can range from minuscule to major -- from a discrete mutation in a single base in the DNA of a single gene to a gross chromosome abnormality involving the addition or subtraction of an entire chromosome or set of chromosomes. Some genetic disorders are inherited from the parents, while other genetic diseases are caused by acquired changes or mutations in a preexisting gene or group of genes. Mutations can occur either randomly or due to some environmental exposure.

There are a number of different types of genetic inheritance, including the following four modes:

Single gene inheritance, also called Mendelian or monogenetic inheritance. This type of inheritance is caused by changes or mutations that occur in the DNA sequence of a single gene. There are more than 6,000 known single-gene disorders, which occur in about 1 out of every 200 births. These disorders are known as monogenetic disorders (disorders of a single gene).

Some examples of monogenetic disorders include:

Single-gene disorders are inherited in recognizable patterns: autosomal dominant, autosomal recessive, and X-linked.

Medically Reviewed by a Doctor on 5/21/2015

Genetic Disease - Symptoms Question: What were the symptoms of a genetic disease in you or a relative?

Genetic Disease - Screening Question: Have you been screened for a genetic disease? Please share your story.

Genetic Disease - Personal Experience Question: Is there a genetic disease in your family? Please share your experience.

Original post:
Genetic Disease: Get the Definition of These Disorders

Read More...

Page 44«..1020..43444546..50..»


2024 © StemCell Therapy is proudly powered by WordPress
Entries (RSS) Comments (RSS) | Violinesth by Patrick