12
171 Molecular Diagnostics Copyright © 2010 Elsevier Ltd. All rights reserved. Detection of Genomic Duplications and Deletions Graham R. Taylor 1 and Carol A. Delaney 2 1 RegionalGeneticsLaboratoryandCR-UKMutationDetectionFacility,StJames’sUniversityHospital, DNALaboratory,Leeds,YorkshireLS97TF,UK; 2 SEScotlandCytogeneticsLaboratory,WesternGeneralHospital,CreweRoad,EdinburghEH42XU,UK Chapter 12 12.1  INTRODUCTION Gene and chromosome duplications have been implicated as fundamental evolutionary mechanisms (Ohno etal.,1968; Li and Gojobori, 1983; Schughart etal., 1989). For the pur- pose of this chapter, duplications and deletions are defined as those too large to be routinely detected by sequencing or simple PCR-based methods (100 bp), but too small to be detected by conventional cytogenetic metaphase analysis (106 kb). Unlike other types of rearrangement, for exam- ple LINE insertions or balanced translocations, a duplica- tion or deletion will be associated with a direct change in gene dosage. This four-order size range includes small frag- ments of genes, exons, entire genes, and multiple genes, and the same methods can, of course, be used to measure changes in chromosome count. Historically, supernumerary chromosomes are the earli- est ‘‘gene’’ duplications reported, revealed by cytogenetic techniques (Lejeune, 1960; Patau et al., 1961). Soon after- ward, evidence for pathogenic -globin gene deletions was reported (Ose and Bush, 1962), although more than 10 years elapsed before direct confirmation was achieved (Ottolenghi et al., 1974). -Globin gene deletions were reported soon afterward (Kan et al., 1975), followed by gene deletions in other hemoglobinopathies and culminating in the use of Southern blotting for the prenatal diagnosis of globin dele- tions (Orkin et al., 1978). Gene dosage changes were next reported in the immunoglobulin genes (Rabbitts etal., 1980; van Loghem et al., 1980) and the first deletion of a tumor suppressor gene (in retinoblastoma) was reported soon afterward (Junien etal., 1982). The identification of repeti- tive DNA associated with a gene deletion was described in a form of hereditary persistence of fetal hemoglobin (Jagadeeswaran etal., 1982). A pathogenic role was attrib- uted to repetitive DNA elements by Hess and coworkers (1983), who suggested that DNA insertion elements may disrupt gene correction processes in the two duplication units containing HBA2 and HBA1 genes. Although the widespread use of Southern blotting (Southern, 1975) from the mid-1970s until the late 1980s may have facilitated the detection of deletions and duplications, the application of the polymerase chain reaction (PCR; Mullis et al., 1986) may unintentionally have produced an ascertainment bias away from them, the analysis of dystrophin gene deletions in males being a notable exception (Beggs et al., 1990). This is because a typical PCR is not designed for quantita- tive analysis, but for optimal purity and yield. However, the observation of germ-line deletions in a wide range of genetic conditions has required the development of techniques that can detect gene dosage changes in hemizygotes. These tech- niques include Southern blotting and PCR modifications, as well as newer methods. Techniques that measure gene dos- age can also be adapted to quantify the somatic mosaicism and PCR failure (allele dropout, see also Chapter 33). 12.2  MECHANISMS Deletions and duplications can be mediated by homolo- gous recombination involving recombinogenic elements, for example PMP22 (Inoue et al., 2001); or gene duplica- tions, for example the type IV collagen genes COL4A5 and COL4A6, paired head to head on chromosome Xq22 dele- tions in X-linked Alport syndrome. BRCA1 gene deletions involving a head-to-head of a partial pseudogene have been reported (Brown etal., 2002), as well as Alu-mediated intra- genic deletions (Puget etal., 1997; Rohlfs etal., 2000). A 26 bp core sequence in two out of five 0 -thalassemia dele- tions has been reported (Harteveld etal., 1997), supporting the idea that Alu repeats stimulate recombination events not

Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

Detection of Genomic Duplications and Deletions

Graham R. Taylor1 and Carol A. Delaney2

1�Regional�Genetics�Laboratory�and�CR-UK�Mutation�Detection�Facility,�St�James’s�University�Hospital,��

DNA�Laboratory,�Leeds,�Yorkshire�LS9�7TF,�UK;2SE�Scotland�Cytogenetics�Laboratory,�Western�General�Hospital,�Crewe�Road,�Edinburgh�EH4�2XU,�UK

Chapter 12

Molecular DiagnosticsCopyright © 2010 Elsevier Ltd. All rights reserved.

12.1  IntroductIon

Gene and chromosome duplications have been implicated as fundamental evolutionary mechanisms (Ohno et�al.,1968; Li and Gojobori, 1983; Schughart et�al., 1989). For the pur-pose of this chapter, duplications and deletions are defined as those too large to be routinely detected by sequencing or simple PCR-based methods (100 bp), but too small to be detected by conventional cytogenetic metaphase analysis (106 kb). Unlike other types of rearrangement, for exam-ple LINE insertions or balanced translocations, a duplica-tion or deletion will be associated with a direct change in gene dosage. This four-order size range includes small frag-ments of genes, exons, entire genes, and multiple genes, and the same methods can, of course, be used to measure changes in chromosome count.

Historically, supernumerary chromosomes are the earli-est ‘‘gene’’ duplications reported, revealed by cytogenetic techniques (Lejeune, 1960; Patau et�al., 1961). Soon after-ward, evidence for pathogenic -globin gene deletions was reported (Ose and Bush, 1962), although more than 10 years elapsed before direct confirmation was achieved (Ottolenghi et� al., 1974). -Globin gene deletions were reported soon afterward (Kan et� al., 1975), followed by gene deletions in other hemoglobinopathies and culminating in the use of Southern blotting for the prenatal diagnosis of globin dele-tions (Orkin et�al., 1978). Gene dosage changes were next reported in the immunoglobulin genes (Rabbitts et�al., 1980; van Loghem et�al., 1980) and the first deletion of a tumor suppressor gene (in retinoblastoma) was reported soon afterward (Junien et�al., 1982). The identification of repeti-tive DNA associated with a gene deletion was described in a form of hereditary persistence of fetal hemoglobin (Jagadeeswaran et�al., 1982). A pathogenic role was attrib-uted to repetitive DNA elements by Hess and coworkers

171

(1983), who suggested that DNA insertion elements may disrupt gene correction processes in the two duplication units containing HBA2 and HBA1 genes. Although the widespread use of Southern blotting (Southern, 1975) from the mid-1970s until the late 1980s may have facilitated the detection of deletions and duplications, the application of the polymerase chain reaction (PCR; Mullis et� al., 1986) may unintentionally have produced an ascertainment bias away from them, the analysis of dystrophin gene deletions in males being a notable exception (Beggs et� al., 1990). This is because a typical PCR is not designed for quantita-tive analysis, but for optimal purity and yield. However, the observation of germ-line deletions in a wide range of genetic conditions has required the development of techniques that can detect gene dosage changes in hemizygotes. These tech-niques include Southern blotting and PCR modifications, as well as newer methods. Techniques that measure gene dos-age can also be adapted to quantify the somatic mosaicism and PCR failure (allele dropout, see also Chapter 33).

12.2  MechanIsMs

Deletions and duplications can be mediated by homolo-gous recombination involving recombinogenic elements, for example PMP22 (Inoue et�al., 2001); or gene duplica-tions, for example the type IV collagen genes COL4A5 and COL4A6, paired head to head on chromosome Xq22 dele-tions in X-linked Alport syndrome. BRCA1 gene deletions involving a head-to-head of a partial pseudogene have been reported (Brown et�al., 2002), as well as Alu-mediated intra-genic deletions (Puget et�al., 1997; Rohlfs et�al., 2000). A 26 bp core sequence in two out of five 0-thalassemia dele-tions has been reported (Harteveld et�al., 1997), supporting the idea that Alu repeats stimulate recombination events not

Page 2: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

Molecular Diagnostics172

only by homologous pairing, but also by providing bind-ing sites for recombinogenic proteins. Deletion of subtelo-meric repeats has been implicated in facioscapulohumeral muscular dystrophy (FSHD) where a remnant fragment, the result of a deletion of tandemly arrayed 3.3 kb repeat units (D4Z4) on 4q35 can be detected (Lemmers et�al., 1998). Non-homologous recombination also has been implicated in gene deletions (Hu and Worton, 1992; Suminaga et�al., 2000).

12.3  PathologIcal consequences

12.3.1  supernumerary chromosomes

In practice, when considering alterations in autosomal chromosome copy number, only supernumerary chromo-somes need to be addressed, as autosomal monosomy is not compatible with life (unless the individual has monosomy mosaicism). Monosomy of the X chromosome can occur (Turner syndrome), although it is estimated that only 1 in 300 conceptuses with monosomy X survive (Kajii et�al., 1980). Triploidy and trisomies of other autosomes are also a common cause of spontaneous abortion. From a clini-cal perspective, prenatal detection of triploidy and triso-mies 13 and 18 are important as fetuses may survive until birth, but most die soon after due to congenital malforma-tions. Trisomies that have less severe consequences are trisomy 21 (Down syndrome), trisomy X (rarely show any physical abnormalities), and Klinefelter syndrome (XXY). Supernumerary chromosomes are readily detectable by conventional cytogenetics (see also Chapter 10); however, molecular techniques may be faster and less expensive.

12.3.2  Microdeletions

Pathogenic partial deletions and duplications have been widely reported for all chromosomes, and cause a range of symp-toms. Some deletions are easily detectable by conventional cytogenetics, others are smaller (microdeletions) and can be difficult or impossible to detect by conventional cytogenetics, and other cytogenetic techniques such as FISH or molecular techniques are required. Some of the more common microde-letions with associated syndromes (summarized in Table 12.1) are 22q11.2 (Di George syndrome), 15q11.2/q12 (Prader-Willi and Angelman syndromes), 17p (Miller-Dieker and Smith-Magenis syndromes), 4p16.3 (Wolf-Hirschhorn syndrome), and 5p15 (Cri du chat syndrome).

12.3.3  subtelomeric deletions

Several syndromes that are caused by microscopically visible chromosomal deletions and duplications, includ-ing the subtelomeric region, have been known to cause idiopathic mental retardation (Flint et� al., 1995; Knight

et� al., 1999) (e.g. 1p, 4p, 5p, 9p). Despite their clinical relevance, screening for more cryptic alterations in the subtelomeres cannot be readily observed by G-banding analysis since most terminal bands are G-band negative. More recent studies using molecular methods such as detection of loss of hypervariable DNA polymorphisms or microsatellite markers and multiprobe FISH that detect submicroscopic subtelomeric deletions have shown that these can account for up to 5% of cases with mental retar-dation (Flint et�al., 1995; Knight et�al., 1999; Rio et�al., 2002). Subtelomeric deletion analysis of all chromosomes can be performed rapidly and cost effectively using new molecular techniques such as MAPH (Sismani et� al., 2001), MLPA (Rooms et� al., 2006; Palomares et� al., 2006), and array CGH (de Vries et�al., 2005).

12.3.4  gene deletions or duplications

A proportion of some types of cancers are caused by inherited germ-line mutations in tumor suppressor genes, and full or partial gene deletions account for a significant number of these mutations. There are a range of other sin-gle gene disorders where deletion or duplication of part or all of the gene accounts for a significant proportion of detected mutations, such as Duchenne muscular dystrophy, spinal muscular atrophy, Charcot-Marie-Tooth disease,

Table 12.1 Proportion (%) of mutations due to deletions or insertions in various genes.

Gene Disorder Deletions or duplications (%)

References

Familial breast cancer

4–27 Puget et al. (1999); Gad et al. (2002); Hogervorst et al. (2003)

VHL Von Hippel- Lindau disease

20–47 Shuin et al. (1995); Vortmeyer et al. (2002)

DMD Duchenne muscular dystrophy

60 Koenig et al. (1989)

MSH2/MLH1 Hereditary non-polyposis colorectal cancer

27–54.8 Wijnen et al. (1998); Gille et al. (2002); Wagner et al. (2002)

RB1 Retinoblastoma 14 Richter et al. (2003)

Page 3: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

chapter  |  12 Detection of Genomic Duplications and Deletions 173

Fanconi anemia, congenital adrenal hyperplasia, and rare metabolic disorders such as non-ketotic hyperglycinemia. Some examples of genes for which an estimate has been obtained for the proportion of germ-line mutations that are deletions or duplications are shown in Table 12.1.

Since the majority of mutations identified in the tumor suppressor genes lead to the production of a truncated prod-uct (Couch and Weber, 1996), until recently most studies performed mutation analysis on genomic DNA using PCR-based techniques such as sequencing, heteroduplex analy-sis (Chapter 4), or the protein truncation test (PTT; Chapter 19). Deletions (Petrij-Bosch et� al., 1997; Puget et�al., 1997; Swensen et� al., 1997) or duplications (Puget et�al., 1999) within the BRCA1 gene, for example, would have been missed by conventional PCR-based methods and were detected by either reverse transcriptase PCR (RT-PCR) or Southern blotting. Although both PCR and Southern blot-ting have been adapted to provide quantitative data, PCR has become the method of choice for genetic testing. Estimates of gene dosage typically have been based on comparisons with a reference standard. Other approaches, including the study of junction fragments or microsatellite inheritance and more recently long-range PCR (Coulter-Mackie et� al., 1998), FISH (Voskova-Goldman et� al., 1997) and array-CGH (Bruder et�al., 2001), have also been employed.

The overall contribution of deletions and duplications recorded in the May 2000 Human Gene Mutation data-base was 5.5% of reported mutations. As of March 2008 the proportion had increased to 22.7%, comprising 16% deletions and 6.5% duplications (http://www.hgmd.cf.ac.uk; Stenson et�al., 2003). Given the greater technical dif-ficulties in identifying deletions and duplications, this is still likely to be an underestimate. This emphasizes the importance of including the measurement of gene dosage in any comprehensive mutation scan. Numerous polymor-phisms have been described involving deletion or duplica-tion of large chromosomal segments, sometimes involving entire genes; for example, the common deletion polymor-phisms of the cytochrome P450 gene CYP2D6 (Meyer and Zanger, 1997), the theta-class glutathione S-transferase gene GSTT1 (Wiencke et�al., 1995), and the mu-type gene GSTM1 (Brockmoller et�al., 1992).

Regions close to telomeres are especially prone to inter-chromosomal rearrangements that can lead to different forms of presence/absence polymorphism, including the multi-allelic variation in the structure of the 16p telomere (Wilkie et�al., 1991), and the deletion polymorphism near the 12q tel-omere (Baird et�al., 2000). The assembly of the human genome sequence is complicated by the presence of duplication/deletion polymorphisms, either because only one form of a region of variable structure is recorded in the sequenced chromosome (Siniscalco et� al., 2000) or because the presence of unsus-pected polymorphism for tandem duplication can lead to missassembly (Bailey et�al., 2001, 2002).

12.4  dIagnostIc technIques

Classical cytogenetic genetic techniques have been valu-able in identifying supernumerary chromosomes and large deletions. However, the resolution of these methods is lim-ited to several megabases and they would miss many known submicroscopic deletions. Fluoresence in�situ hybridization (FISH) techniques have greatly expanded the capabilities of cytogenetics (see also Chapter 10); single copy probes allow detection of microdeletions, which would be impos-sible to detect by conventional cytogenetics; however, due to the specialized equipment and expertise required and the low throughput, detection of submicroscopic deletions has largely become the realm of specialized molecular genetic diagnostics. The detection of submicroscopic duplications by cytogenetics remains a challenge; the use of interphase FISH together with image analysis is being used increasingly.

12.4.1  southern Blotting

Southern blotting still has a role in a comprehensive diagnos-tic service, both for gene dosage measurement and for the measurement of allele expansions. Southern blots can detect deletions or duplications by the identification of novel restric-tion fragments created by the rearrangement or by measure-ment of band intensity compared to a control fragment.

Fragment size changes are seen because the region of genomic DNA created by a deletion or duplication may have gained or lost restriction enzyme sites. In those cases, a Southern blot of a genomic digest using an appropriate restriction enzyme will detect novel fragments. There is a risk that some altered fragment sizes may simply be due to restriction fragment length polymorphisms, although the risk can be reduced by performing separate digestions with a different enzyme. Direct visualization of fragment sizes has long been used for detecting -globin duplications and deletions (Orkin et�al., 1978), and the increased size range available in pulse field gel electrophoresis (PFGE) allows detection of deletions and duplications even in genes as large as the dystrophin gene (Kenwrick et�al., 1987).

Estimation of gene dosage by measuring the intensity of probe hybridization (usually in comparison with a control) has identified several instances of gene deletions (Bonifas and Epstein, 1990), though few studies report details of the dose–response curve (such as its linear range) and inter-nal controls are often lacking. Dystrophin deletion carrier testing is possible by assessing relative band intensity (van Essen et�al., 1997). Accurate measurement of band inten-sity requires a phosphoimager, which gives a linear dose response over a wider range than film emulsion. Figure 12.1 shows the application of Southern blotting to measure the relative gene dosage of CYP21A and CYP21B, which can be deleted in congenital adrenal hyperplasia. The risk of inaccuracies introduced by uneven transfer of DNA to the membrane or incomplete washing of the probe requires

Page 4: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

Molecular Diagnostics174

that an additional probe for a control locus should be included as a standard.

12.4.2  Microsatellites and snPs

Microsatellites have been used to detect supernumerary chromosomes (Mansfield, 1993) and large deletions or duplications (Brice et� al., 1992). This approach is limited by the fact that semi-quantitative PCR methods do not work well using microsatellites, so only cases in which the allele lengths of the microsatellites differ over chromosomal region of interest can be interpreted unambiguously. Peak heights or areas of microsatellites can be hard to interpret because of preferential amplification of the smaller allele and the stuttering effect that produces a series of minor peaks imme-diately adjacent to the major peak. The difficulties with mul-tiplex quantitative fluorescent PCR (QF-PCR) are reduced by selection of tetranucleotide repeat markers and is finding increasing use in the rapid diagnosis of common aneuploi-dies (Donaghue et�al., 2003). Loss of heterozygosity of sin-gle nucleotide polymorphisms (SNPs) using high-throughput Affymetrix human SNP arrays has been proposed as a rapid means of identifying allele imbalance caused by genomic deletions in tumor cells (Hoque et�al., 2003).

12.4.3  Polymerase chain reaction

Neubauer and colleagues (1990) described the co-amplification of two amplicons followed by quantitation on the basis of staining intensity as a means of estimating gene dosage.

FIgure 12.1  Gene dosage in 21 hydroxylase by Southern blotting. The Southern blot shows four major bands: C4A (7 kb), C4B (6 or 5.4 kb), 21-hydroxylase (3.7 kb), and 21-hydroxylase pseudogene (indicated by an asterisk). Sample 1 has a homozygous deletion of 21-OH, sample 2 has equal gene dosage at all four loci. Samples 3 and 4 each have deletions of the 21-hydroxylase pseudogene and C4A, and sample 5 has a deletion of 21-hydroxylase and C4B. (Image courtesy of Dr. Kieran Bransfield, Leeds Institute of Molecular Medicine, Leeds.)

Differential PCR-based methods are semi-quantitative: they determine relative concentrations of two amplicons, but not absolute molar amounts. Taking advantage of the increased sensitivity of fluorescent detection methods Yau and coworkers (1996) described a multiplex fluorescent PCR that was able to detect female carriers of deletions or duplications in the dystrophin gene. Using peak areas, they were able to give statistical estimates of their assay, likely to be essential for diagnostic applications. Similar meth-ods have been used for testing for hereditary motor and sensory neuropathy (HMSN) duplications (Rowland et� al., 2001) and APC deletions (Flintoff et�al., 2001). Casilli and coworkers (2002) refined the basic multiplex fluorescent PCR method by selecting shorter fragments, tagging the primers with common tags of 16 nucleotides, and using a modified PCR buffer containing DMSO. The method was termed quantitative multiplex PCR of short fluorescent frag-ments (QMSFPCR) and enables rapid design of amplicons to fine-map the limits of a deletion or duplication (Casilli et�al., 2002). Although differential PCR methods offer high orders of multiplicity (e.g. 8–15 amplicons per reaction), end-point multiplex PCR assays rely on the equivalence of amplifica-tion of each fragment in the multiplex in the test sample as well as a control sample. This equivalence may be lost if the starting template DNA concentration is too variable, since different fragments within the multiplex may be amplified with different efficiencies. In some situations it may be nec-essary to choose a more robust, lower order multiplex base on real-time PCR. Real-time PCR (also known as quantita-tive PCR; qPCR) provides a means for continuous detection and quantification of product throughout the amplification process, and as such can dispense with a gel separation stage and operate in a closed system (see also Chapter 7). The accumulation of PCR product is monitored by stain-ing using interchelating dyes (e.g. SYBR Green) or by dual-labeled probes such as TaqMan (Laurendeau et�al., 1999), molecular beacons (Tyagi and Kramer, 1996), or other fluorescent detection systems. Real-time PCR is becom-ing widely used as a method for measuring gene dosage (Feldkotter et�al., 2002; Bertin et�al., 2003; Covault et�al., 2003; Gaikovitch et�al., 2003; Kim et�al., 2003). Real-time quantitative PCR has been used for screening large gene and chromosomal rearrangements (Ariani et� al., 2004). Additionally, real-time PCR can increase the sensitivity of mutation detection especially in double-copy genes such as X-linked methyl-CpG-binding protein 2 (MECP2) gene, where current mutation scanning techniques such as DGGE, SSCP, DHPLC and direct sequencing are prone to miss gross rearrangements.

12.4.4  MaPh and MlPa

The multiplex amplifiable probe hybridization (MAPH) method for copy-number measurement (Armour et� al., 2000) combines hybridization as the primary step to detect copy number with end-point multiplex PCR to amplify the

Page 5: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

chapter  |  12 Detection of Genomic Duplications and Deletions 175

hybridized probes. Sets of short probes corresponding to the segments to be tested, each flanked by the same primer-binding sites, are hybridized with the test genomic DNA immobilized on a solid support. After washing, each spe-cifically bound probe will be present in an amount propor-tional to its copy number. All probes can then be amplified simultaneously with a single primer pair, and quantified after electrophoretic separation. This study demonstrated the simultaneous assessment of copy number in a set of 40 human loci, including detection of deletions causing Duchenne muscular dystrophy and Prader-Willi/Angelman syndromes. The high order multiplex achieved may be lim-ited only by the need to have probes of varying size for electrophoretic separation. MAPH probes are generated by cloning the target sequences into a plasmid vector, amplify-ing the cloned sequence using primers directed to the vec-tor with the result that all probes are then flanked with the same sequence. Probes that are intended to be multiplexed must be of sufficient size difference to be resolved by elec-trophoresis. The membranes are then washed rigorously to remove unbound probe, and the remaining specifically bound probe will be present in an amount proportional to its target copy number. The probes are then stripped from the membrane by boiling, and amplified simultaneously with the universal primer pair. Products are then sepa-rated by electrophoresis, and a relative comparison is made between the peak heights. Reduced peak heights compared to internal control probes indicate a reduction in gene copy number (deletion) and an increase in gene copy number (duplication) produces increased peak heights. The assay can be completed in two to three days, requiring one over-night hybridization followed by membrane washing, a PCR step, and a product detection step. The system works well, but the manipulation of small nylon filters presents some difficulties in sample handling and labeling.

A similar technique called multiplex ligatable probe amplification (MLPA) avoids the use of filters by using single-stranded ligatable probes and a thermo-stable DNA ligase to produce the amplifiable target (Schouten et� al., 2002; see also Chapter 13). In the MLPA technique, genomic DNA is hybridized in solution to probe sets, each of which consists of two halves. One half consists of a target- specific sequence (20–30 nucleotides) flanked by a universal primer sequence, and can be generated synthetically. The other half also has a target-specific sequence at one end (25–43 nucleotides) and a universal primer sequence at the other, but has a variable length stuffer fragment in between (19–370 nucleotides) to generate the size differences necessary in the probes to allow electrophoretic resolu-tion. This larger probe part is generated by cloning the target-specific sequence into M13 derived vectors that already contain the variable length fragments; single-stranded DNA is then purified from the phage particles and made double stranded at two sites by annealing short oligonucle-otides in order that the desired probe fragment be liberated by restriction enzyme digestion. The two probe halves are

designed such that the target-specific sequences bind adja-cently to the target DNA, and can then be joined by use of a ligase. This generates a contiguous probe flanked by universal primer binding sites that can then be amplified by PCR, whereas unbound probe halves cannot be amplified, and hence eliminates the need for removal of excess probe by washing. The amounts of ligated probe produced will be proportional to the target copy number, and after PCR amplification the relative peak heights indicate deletion or duplication of target sequence.

In diagnostic use, applications of these high order mul-tiplex methods include the detection of aneuploidies, unbal-anced cryptic translocations, and whole or partial gene duplications or deletions. Commercial kits based on MAPH and MLPA are available (http://www.mrc-holland.com). The high multiplicity of MAPH and MLPA, plus their use of standard genetic laboratory apparatus, make it highly likely that these techniques will become widely used. Both techniques offer a rapid means of scanning up to 40 loci for gene dosage, and are likely to be used widely in research and diagnostic settings. MAPH represents a conceptual breakthrough for the analysis of gene dosage, but the han-dling of small filter discs is difficult and could pose sample tracking problems in routine medium throughput settings. The liquid-phase solution to sample handling offered by MLPA and the ready availability of robust commercial kits has led to rapid acceptance by diagnostic laboratories worldwide. MLPA uses M13 (single-stranded) probes that are more technically challenging to construct than MAPH probes, which can be made from PCR products.

There is an ever-growing range of MLPA probe sets and kits supplied by MRC-Holland. However, it is also possible to construct additional probes to enable fine mapping and identification of the junction fragments/deletion breakpoints by designing your own synthetic MLPA probes. The advan-tage of synthetic probes, as compared to the phage M13-derived probes made by MRC-Holland, is that the cloning step is omitted, so they are available sooner, but disadvan-tages are that fewer probes can be used in a single MLPA reaction, approximately 11 bespoke probes can be added to a probemix. Protocols assisting design of synthetic probes are available from the MRC-Holland website or using AlleleID probe design software (http://www.premierbiosoft.com/datafiles/alleleIDWin.exe). Alternatively, additional loci could be examined by adapting MLPA to an array detection setting, so that fragments were identified by sequence rather than size. This would enable all MLPA probes to be chemi-cally synthesized. Array adaptations of both MAPH and MLPA might enable much higher order multiplexes than are currently possible. However, even 40-plex assays represent a substantial gain in the multiplex order typically available in PCR. MAPH or MLPA could provide competition for the array-based CGH approaches: a 96-well array of 40 probes could interrogate over 3,000 loci, representing a better than 1 cM coverage of the entire human genome or much higher single chromosome resolution. Two color MLPA kits are

Page 6: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

Molecular Diagnostics176

available increasing the number of loci examined to some extent by including two probe sets which are amplified by two sets of primers: one labeled with FAM and the other with HEX. An example is the MLPA kit to detect deletions and duplications in the CBP (Crebb binding protein) gene. In all, this kit has synthetic probes for 20 out of 32 exons of CBP, and also five control probes for unlinked loci. The probes span the CBP gene from exons 2 to 32 (the coding region of the gene) and, due to the number of probes, is able to detect deletions and duplications in CBP in cases with Rubinstein-Taybi syndrome previously undetectable by FISH (Roelfsema et�al., 2005).

12.4.5  long-range Pcr

Long-range PCR (Barnes, 1992) is a modified PCR protocol that includes a proof-reading polymerase and short denatur-ation times to enable the size of amplicons to be increased from 3–5 kb to beyond 30 kb. This opens the possibility of screening for deletions of this scale by direct PCR, with the advantage of producing the junction fragment for further analysis. Long-range PCR has been used successfully to identify deletions including mitochondrial DNA (Fromenty et�al., 1996), C4 gene deletions in the MHC complex (Grant et� al., 2000), the CYP2D6 deletion allele, and LDL dele-tions (Kim et�al., 1999). Despite the simplicity in principle of a long-range PCR approach, in practice it is often dif-ficult to design a robust assay. Direct detection of duplica-tions by amplification of entire duplicated regions is likely to prove difficult in diploid genomes since the unduplicated allele will have a significant advantage during the amplifi-cation process. However, full characterization of a deletion or duplication requires the sequence of the junction frag-ments. This can be achieved either by using long-range PCR or by fine mapping, for example using QMSF-PCR followed by conventional PCR and sequencing.

12.4.6  array-Based cgh

Microarrays have become widely used tools for gene expression studies, and now encouraging developments are taking place in array-based comparative genome hybridiza-tion (Array-CGH) as a tool for measuring gene duplications and deletions. Comparative genomic hybridization origi-nally used metaphase chromosomes as targets for differen-tially labeled probes (e.g. Cy5, Cy3) from control and test samples. Gene dosage changes could be detected by varia-tion of the relative intensity of the two labels (Kallioniemi et� al., 1992). By replacing the metaphase spread with microarrayed BAC DNA, Pinkel and colleagues (1998) reported the use of array-based CGH analysis to investigate chromosome 20 gene dosage alterations in breast cancer. DNA purified from BAC clones spaced at approximately 3 Mb intervals along the entire chromosome together with

some X chromosome controls were arrayed on nitrocel-lulose membranes as 200–400 micron spots in duplicate. Test DNA from breast cancer cell-lines was labeled by nick translation with fluorescein, control DNA with Texas Red, and the spots were counterstained with DAPI. Scanning used custom-built mercury-arc illumination and CCD detection. Other studies have used different dyes (with Cy3/Cy5 being a popular choice) and omitted counterstain-ing. Many detection systems use confocal laser scanners, which, although more restricted in fluorochrome options, are brighter and have fewer problems with light scattering.

Growing high numbers (104) of cloned DNA are expensive and labor intensive. To circumvent large-scale cultures and DNA purifications, array features have been made using degenerate oligonucleotide PCR (DOP-PCR) of BAC templates using 5 amine-linked primers (Hodgson et� al., 2001). Fiegler and coworkers (2003) improved the standard DOP-PCR to reduce non-specific host and vec-tor amplicons from BAC clones, improving the signal-to-noise ratio. A ligation-mediated PCR BAC labeling method has also been reported that improved signal-to-noise ratios (Snijders et�al., 2001). Array CGH is still in development (see also Chapters 10 and 12), but strong commercial and academic interest are likely to result in an increase in research and diagnostic applications in the near future. The supply of robust competitor (Cot1) DNA has been a major problem reported by many array-CGH users (Carter et�al., 2002). Cot1 DNA is used to block the hybridization of repetitive DNA in BAC-derived amplicons. Buckley and coworkers (2002) constructed a comprehensive microarray representing a human chromosome for analysis of DNA copy-number variation. The chromosome 22 microarray covered 34.7 Mb with an average resolution of 75 kb using a sequence-defined, repeat-free, and non-redundant strat-egy for array preparation. This enabled an increase in array resolution and eliminated the need for Cot1 DNA. Array targets were made using phi29 DNA polymerase synthesis.

As the technology becomes more widely used, com-mercial suppliers are starting to emerge; for example, the Cytochip array from BlueGnome (http://www.cytochip.com) is designed for diagnostic use. This is a BAC micro-array with an average genome coverage of approximately 565 kb, with median coverage of 250 kb in subtelomeric regions and an average resolution of 100 kb in 90 known clinically relevant regions (microdeletion/duplication syn-dromes). Higher resolution oligonucleotide-based microar-rays for the detection of copy-number changes are becoming increasingly popular allowing genome-wide analysis which can map chromosomal imbalance breakpoints at exon level resolution, including imbalances that are single copy- number genomic alterations (Selzer et�al., 2005). Commercial suppliers of oligonucleotide-based arrays include Agilent Technologies Inc. (http://www.agilent.com). Several Agilent array CGH platforms exist including arrays with 44,000 60-mer oligonucleotide probes covering both coding and

Page 7: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

chapter  |  12 Detection of Genomic Duplications and Deletions 177

non-coding human sequences with an average genome res-olution of 50 kb through to 244,000 probes with a mediam probe spacing of 9 kb. Other suppliers of high-resolution oligonucleotide-based arrays to detect gene dosage changes include Illumina (http://www.illumina.com) and Affymetrix (http://www.affymetrix.com). Additionally, genome centers, for example, Leiden University in Holland and the Sanger Centre in the UK, are using their resources to manufacture genome- or chromosome-specific arrays.

Since array-CGH enables high-resolution detection of structural variation, this technique is moving from a research tool into diagnostic service provision. Confirmation of the array CGH results are often needed prior to issu-ing a diagnostic report. Several copy number/dosage tech-niques confirm array CGH results, including FISH, MLPA, STR analysis, real-time PCR/qPCR, long-range PCR, and, ultimately, DNA sequencing. The choice of confirmation method will depend upon several factors, including the size of the region involved, the precise location of the abnormal-ity, availability of FISH or MLPA probes, the density of STRs, the material available from the proband and the par-ents (DNA, metaphase spreads, etc.), and access to equip-ment. Often there is an interplay between several techniques to determine copy-number variation per case (Fig. 12.2).

One of the findings of using higher-resolution array CGH for genomic-wide copy-number changes is the awareness of the significant degree of natural copy-number variation in the genome. At 100 kb array CGH resolution Redon and coworkers (2006) reported an average of 70 polymorphisms per patient and a total of 1,116 unique polymorphisms on over 270 patients. Similarly, de Vries and coworkers (2005) reported 4–9 copy-number variants per patient in a cohort of mental retardation patients, where 96% of these variations were found to be inherited polymorphisms.

12.5  noMenclature

Recently, the Human Genome Variation Society (http://www.hgvs.org) has taken on the responsibility of establishing a standardized mutation nomenclature that includes deletions and duplications. A nomenclature reference is maintained at this site by Dr. Johan Den Dunnen, Leiden University Medical Centre. The current recommendation for exon or multi-exon deletion nomenclature varies, depending on whether or not the breakpoint has been identified. If the breakpoints are not sequenced (e.g. detected on Southern blot or by MLPA), exonic deletions are described as c.88-?_923?del, indicating a deletion starting at an unknown posi-tion in the intron 5 of cDNA nucleotide 88 and ending at an unknown position in the intron 3 of cDNA nucleotide 923. A genomic or cDNA reference (e.g. Genbank, EMBL, DDJB) should be cited, including the version number. If the junc-tion fragments are known, then the form g.390_1458del or (g.390_1458del1069) should be used for a genomic reference

sequence or c.13-23_301-143del (c.13-23_301-143del1069) for a cDNA reference sequence.

12.6  gene dosage aPPlIcatIons In tuMor ProFIlIng

Partial chromosomal losses occur relatively frequently in a large number of tumor types, and this is readily detectable by demonstrating loss of heterozygosity (LOH) of polymorphic

FIgure 12.2  Detection of a duplication of 16qter in a patient with unexplained learning difficulties using: A. A 1 Mb BAC array from BlueGnome. B. The MLPA subtelomere kit from MRC-Holland confirms the duplication, and C. Interphase FISH using a Vysis subtelomeric probe. Duplication of 16qter is shown by three copies of the red signal as indicated by arrows. Note: The e-book for this title, including full-color images, is available for purchase at www.elsevierdirect.com.

Page 8: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

Molecular Diagnostics178

microsatellite markers. Such techniques have been widely employed to investigate various aspects of tumor development:

1.  Clonality. Loss of chromosomal regions is essentially an irreversible event in tumor cell development, and all subsequent subclones would be expected to demon-strate the same deletions, and may have accumulated more. LOH analysis of multiple tumors from the same patient can be used to determine whether tumors, either synchronous or metachronous, are clonal in origin, and in the latter case may be used to determine whether subsequent tumors are recurrences or metastases of previous tumors or independent primary tumors. For example, in patients with multiple synchronous lung tumors it is important for treatment decisions to discriminate multicentric lung cancers from intrapul-monary metastases, and this can be aided by clonality studies (Shimizu et�al., 2000).

2.  Identification of tumor suppressor genes. It is hypoth-esized that the reason that tumors frequently undergo LOH, and that LOH of particular regions is commonly associated with specific tumor types, is that the regions lost harbor tumor suppressor genes. Identification of such tumor suppressor genes may provide useful tools for diagnosis, prognosis, and possible therapies. Hence, numerous studies have been undertaken to identify regions of LOH specific to individual tumor types and the genes present in these regions. For example, chro-mosome 9 is the most frequently deleted chromosome in transitional cell carcinoma of the bladder, and candidate genes have been identified in three of the four regions of minimal deletion (Knowles, 1999). In prostate can-cer, gains at Xq and 18q are among the most common chromosomal alterations, and amplification of the AR (Xq12), MYC (8q24), and EIF3S3 (8q23) genes have been found in a large fraction of hormone-refractory prostate cancers (Nupponen and Visakorpi, 2000).

3.  Diagnosis. LOH analysis has been employed both in detection of the presence of a tumor, as well as dif-ferential diagnosis and typing of tumors. For example, LOH has been reliably demonstrated in DNA extracted from urine sediments in bladder cancer patients, and may have a role in the non-invasive diagnosis of blad-der cancer (Linn et�al., 1997; Berger et�al., 2002). Differential diagnosis between renal oncocytomas and renal cell carcinomas can be difficult due to morpho-logical similarities, but is important due to their differ-ent prognoses. LOH analysis has shown that they can be differentiated on the basis of spectrum of chromo-somal loss (Herbers et�al., 1998). LOH analysis has also been used to characterize the aggressive intraductal carcinoma of the prostate, to differentiate it from the less aggressive high-grade dysplasia (prostatic intraepi-thelial neoplasia, PIN), and to provide evidence that it does not represent invasion of Gleason grade 3 cancers into the ductal/acinar system (Dawkins et�al., 2000).

4.  Prognosis. In several tumor types, LOH of specific regions have been hypothesized to be key events in tumor evolution and progression. LOH analysis of tumor speci-mens has been used to determine their value as prognostic markers. It has been found in colorectal cancers that high level LOH correlated with earlier onset and lymphatic invasion, and hence a poorer prognosis, and low level LOH was more common in earlier stage disease and pre-dicted a more favorable outcome (Choi et�al., 2002). In breast cancers, specific chromosomal regions have been identified for which LOH is a significant predictor of lymph-node metastasis and hence may serve as a negative prognostic indicator (Nagahata et�al., 2002).

12.7  suMMary and Future develoPMents

Gene dosage is a significant contributor to the overall bur-den of the germ-line and somatic mutations in man. Any comprehensive mutation screen therefore should include measurement of gene dosage. The permeation of genomics into medical practice will increase the demand for mutation screening with applications in diagnosis, predictive testing, and treatment. This, in turn, will encourage the develop-ment of more highly automated approaches, the introduc-tion of robust statistical analysis, and quality control of results that may be used in a diagnostic context. Those techniques are most easily adaptable to robust laboratory processes and automated data handling will have a com-petitive advantage. Ease of handling, hence more robust results, perhaps explains the greater uptake of MLPA com-pared with MAPH. Both MAPH and MLPA produce data using automated DNA sequencers that are spreadsheet-ready. Statistical parameters may then be easily applied as quality measures (Taylor et�al., 2003). The same is true of array scanner output and real-time PCR systems, in con-trast to classical genetic techniques like Southern blotting or cytogenetics. Whereas MLPA, real-time PCR, QMSF-PCR, and MAPH approaches are focused around relatively low numbers of targets, the array methods offer very high orders of multiplicity, the so-called ‘‘hypothesis-free’’ approach. It seems likely that the two approaches will converge as arrays become increasingly targeted toward regions or functional clusters of interest and the PCR-based approaches move toward an array-like format, either in capillary arrays or high density microtiter arrays.

Finally, as clonal sequencing technology matures it may displace the current generation of array technology. This reflects advantages in the clonal/digital nature of the out-put in contrast to the analog output from arrays, the unbi-ased readout, not dictated by what has been arrayed, and the ability to read 30–400 base strings of DNA sequence rather than infer sequence results from hybridization to arrayed sequence of 20 to 80 bases. Structural variations, including copy-number variants, can be readily detected using paired

Page 9: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

chapter  |  12 Detection of Genomic Duplications and Deletions 179

reads, although the sensitivity and specificity of this method has not yet been established. Campbell and coworkers (2008) characterized 306 germ-line structural variants and 103 somatic rearrangements to the base-pair level of resolution.

reFerences

Ariani, F., Mari, F., Pescucci, C., Longo, I., Bruttini, M., Meloni, I., Hayek, G., Rocchi, R., and Zappella, M. (2004). Real time quantita-tive PCR as a routine method for screening large rearrangements in Rett syndrome. Report of one case of MECP2 deletion and one case of MECP2 duplication. Hum.�Mutat. 24, 172–177.

Armour, J.A., Sismani, C., Patsalis, P.C., and Cross, G. (2000). Measurement of locus copy number by hybridisation with amplifiable probes. Nucleic�Acids�Res. 28, 605–609.

Bailey, J.A., Yavor, A.M., Massa, H.F., Trask, B.J., and Eichler, E.E. (2001). Segmental duplications: organization and impact within the current human genome project assembly. Genome�Res. 11, 1005–1017.

Bailey, J.A., Gu, Z., Clark, R.A., Reinert, K., Samonte, R.V., Schwartz, S., Adams, M.D., Myers, E.W., Li, P.W., and Eichler, E.E. (2002). Recent segmental duplications in the human genome. Science 297, 1003–1007.

Baird, D.M., Coleman, J., Rosser, Z.H., and Royle, N.J. (2000). High lev-els of sequence polymorphism and linkage disequilibrium at the tel-omere of 12q: implications for telomere biology and human evolution. Am.�J.�Hum.�Genet. 66, 235–250.

Barnes, W.M. (1992). The fidelity of Taq polymerase catalyzing PCR is improved by an N-terminal deletion. Gene 112, 29–35.

Beggs, A.H., Koenig, M., Boyce, F.M., and Kunkel, L.M. (1990). Detection of 98% of DMD/BMD gene deletions by polymerase chain reaction. Hum.�Genet. 86, 45–48.

Berger, A.P., Parson, W., Stenzl, A., Steiner, H., Bartsch, G., and Klocker, H. (2002). Microsatellite alterations in human bladder cancer: detection of tumor cells in urine sediment and tumor tissue. Eur.�Urol. 41, 532–539.

Bertin, R., Acquaviva, C., Mirebeau, D., Guidal-Giroux, C., Vilmer, E., and Cave, H. (2003). CDKN2A, CDKN2B, and MTAP gene dosage permits precise characterization of mono and bi-allelic 9p21 deletions in childhood acute lymphoblastic leukemia. Genes� Chromosomes�Cancer 37, 44–57.

Bonifas, J.M., and Epstein, E.H., Jr. (1990). Detection of carriers for X-linked ichthyosis by Southern blot analysis and identification of one family with a de novo mutation. J.�Invest.�Dermatol. 95, 16–19.

Brice, A., Ravise, N., Stevanin, G., Gugenheim, M., Bouche, P., Penet, C., and Agid, Y. (1992). Duplication within chromosome 17p11.2 in 12 families of French ancestry with Charcot-Marie-Tooth disease type 1a. The French CMT Research Group. J.�Med.�Genet. 29, 807–812.

Brockmoller, J., Gross, D., Kerb, R., Drakoulis, N., and Roots, I. (1992). Correlation between trans-stilbene oxide-glutathione conjugation activity and the deletion mutation in the glutathione S-transferase class mu gene detected by polymerase chain reaction. Biochem.�Pharmacol. 43, 647–650.

Brown, M.A., Lo, L.J., Catteau, A., Xu, C.F., Lindeman, G.J., Hodgson, S., and Solomon, E. (2002). Germline BRCA1 promoter deletions in UK and Australian familial breast cancer patients: identification of a novel deletion consistent with BRCA1:psiBRCA1 recombination. Hum.�Mutat. 19, 435–442.

Bruder, C.E., Hirvela, C., Tapia-Paez, I., Fransson, I., Segraves, R., Hamilton, G., Zhang, X.X., Evans, D.G., Wallace, A.J., Baser, M.E., Zucman-Rossi, J., Hergersberg, M., Boltshauser, E., Papi, L., Rouleau, G.A., Poptodorov, G., Jordanova, A., Rask-Andersen, H., Kluwe, L.,

Mautner, V., Sainio, M., Hung, G., Mathiesen, T., Moller, C., Pulst, S.M., Harder, H., Heiberg, A., Honda, M., Niimura, M., Sahlen, S., Blennow, E., Albertson, D.G., Pinkel, D., and Dumanski, J.P. (2001). High resolution deletion analysis of constitutional DNA from neurofi-bromatosis type 2 (NF2) patients using microarray-CGH. Hum.�Mol.�Genet. 10, 271–282.

Buckley, P.G., Mantripragada, K.K., Benetkiewicz, M., Tapia- Paez, I., Diaz De Stahl, T., Rosenquist, M., Ali, H., Jarbo, C., De Bustos, C., Hirvela, C., Sinder Wilen, B., Fransson, I., Thyr, C., Johnsson, B.I., Bruder, C.E., Menzel, U., Hergersberg, M., Mandahl, N., Blennow, E., Wedell, A., Beare, D.M., Collins, J.E., Dunham, I., Albertson, D., Pinkel, D., Bastian, B.C., Faruqi, A.F., Lasken, R.S., Ichimura, K., Collins, V.P., and Dumanski, J.P. (2002). A full-coverage, high-resolution human chromosome 22 genomic microarray for clinical and research applications. Hum.�Mol.�Genet. 11, 3221–3229.

Carter, N.P., Fiegler, H., and Piper, J. (2002). Comparative analysis of comparative genomic hybridization microarray technologies: report of a workshop sponsored by the Wellcome Trust. Cytometry 49, 43–48.

Casilli, F., Di Rocco, Z.C., Gad, S., Tournier, I., Stoppa-Lyonnet, D., Frebourg, T., and Tosi, M. (2002). Rapid detection of novel BRCA1 rearrangements in high-risk breast-ovarian cancer families using mul-tiplex PCR of short fluorescent fragments. Hum.�Mutat. 20, 218–226.

Campbell, P.J., Stephens, P.J., Pleasance, E.D., O’Meara, S., Li, H., Santarius, T., Stebbings, L.A., Leroy, C., Edkins, S., Hardy, C., Teague, J.W., Menzies, A., Goodhead, I., Turner, D.J., Clee, C.M., Quail, M.A., Cox, A., Brown, C., Durbin, R., Hurles, M.E., Edwards, P.A., Bignell, G.R., Stratton, M.R., and Futreal, P.A. (2008). Identification of somati-cally acquired rearrangements in cancer using genome-wide massively parallel paired-end sequencing. Nat.�Genet. 40, 722–729.

Choi, S.W., Lee, K.J., Bae, Y.A., Min, K.O., Kwon, M.S., Kim, K.M., and Rhyu, M.G. (2002). Genetic classification of colorectal cancer based on chromosomal loss and microsatellite instability predicts survival. Clin.�Cancer�Res. 8, 2311–2322.

Couch, F.J., and Weber, B.L. (1996). Mutations and polymorphisms in the familial early-onset breast cancer (BRCA1) gene. Breast Cancer Information Core. Hum.�Mutat. 8, 8–18.

Coulter-Mackie, M.B., Applegarth, D.A., Toone, J.R., and Gagnier, L. (1998). A protocol for detection of mitochondrial DNA deletions: characterization of a novel deletion. Clin.�Biochem. 31, 627–632.

Covault, J., Abreu, C., Kranzler, H., and Oncken, C. (2003). Quantitative real-time PCR for gene dosage determinations in microdeletion geno-types. Biotechniques 35, 594–596, 598.

Dawkins, H.J., Sellner, L.N., Turbett, G.R., Thompson, C.A., Redmond, S.L., McNeal, J.E., and Cohen, R.J. (2000). Distinction between intra-ductal carcinoma of the prostate (IDC-P), high-grade dysplasia (PIN), and invasive prostatic adenocarcinoma, using molecular markers of cancer progression. Prostate 44, 265–270.

De Vries, B.B., Pfundt, R., Leisink, M., Kooles, D.A., Lisenka, E.L., Vissers, M., Janssen, I.M., van Reijmarsdal, S., Nillesen, W.M., Huys, E.H., de Leeuw, N., Smeets, D., Sistermans, E.A., Feuth, T., van Ravenswaaijarti, C.M., van Kessel, A.G., Schoenmakers, E.F., Brunner, H.G., and Veltman, J.A. (2005). Diagnostic genome profil-ing in mental retardation. Am.�J.�Hum.�Genet. 77, 606–616.

Donaghue, C., Roberts, A., Mann, K., and Ogilvie, C.M. (2003). Development and targeted application of a rapid QF-PCR test for sex chromosome imbalance. Prenat.�Diagn. 23, 201–210.

Feldkotter, M., Schwarzer, V., Wirth, R., Wienker, T.F., and Wirth, B. (2002). Quantitative analyses of SMN1 and SMN2 based on real-time lightCycler PCR: fast and highly reliable carrier testing and predic-tion of severity of spinal muscular atrophy. Am.� J.�Hum.�Genet. 70, 358–368.

Page 10: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

Molecular Diagnostics180

Fiegler, H., Carr, P., Douglas, E.J., Burford, D.C., Hunt, S., Scott, C.E., Smith, J., Vetrie, D., Gorman, P., Tomlinson, I.P., and Carter, N.P. (2003). DNA microarrays for comparative genomic hybridization based on DOP-PCR amplification of BAC and PAC clones. Genes�Chromosomes�Cancer 36, 361–374.

Flint, J., Wilkie, A.O., Buckle, V.J., Winter, R.M., Holland, A.J., and McDermid, H.E. (1995). The detection of subtelomeric chromo-somal rearrangements in idiopathic mental retardation. Nat.�Genet. 9, 132–140.

Flintoff, K.J., Sheridan, E., Turner, G., Chu, C.E., and Taylor, G.R. (2001). Submicroscopic deletions of the APC gene: a frequent cause of famil-ial adenomatous polyposis that may be overlooked by conventional mutation scanning. J.�Med.�Genet. 38, 129–132.

Fromenty, B., Manfredi, G., Sadlock, J., Zhang, L., King, M.P., and Schon, E.A. (1996). Efficient and specific amplification of identi-fied partial duplications of human mitochondrial DNA by long PCR. Biochim.�Biophys.�Acta. 1308, 222–230.

Gad, S., Caux-Moncoutier, V., Pages-Berhouet, S., Gauthier-Villars, M., Coupier, I., Pujol, P., Frenay, M., Gilbert, B., Maugard, C., Bignon, Y.J., Chevrier, A., Rossi, A., Fricker, J.P., Nguyen, T.D., Demange, L., Aurias, A., Bensimon, A., and Stoppa-Lyonnet, D. (2002). Significant contribution of large BRCA1 gene rearrangements in 120 French breast and ovarian cancer families. Oncogene 21, 6841–6847.

Gaikovitch, E.A., Cascorbi, I., Mrozikiewicz, P.M., Brockmoller, J., Frotschl, R., Kopke, K., Gerloff, T., Chernov, J.N., and Roots, I. (2003). Polymorphisms of drug-metabolizing enzymes CYP2C9, CYP2C19, CYP2D6, CYP1A1, NAT2 and of Pglycoprotein in a Russian population. Eur.�J.�Clin.�Pharmacol. 59, 303–312.

Gille, J.J., Hogervorst, F.B., Pals, G., Wijnen, J., Van Schooten, R.J., Dommering, C.J., Meijer, G.A., Craanen, M.E., Nederlof, P.M., De Jong, D., McElgunn, C.J., Schouten, J.P., and Menko, F.H. (2002). Genomic deletions of MSH2 and MLH1 in colorectal cancer families detected by a novel mutation detection approach. Br.� J.� Cancer 87, 892–897.

Grant, S.F., Kristjansdottir, H., Steinsson, K., Blondal, T., Yuryev, A., Stefansson, K., and Gulcher, J.R. (2000). Long PCR detection of the C4A null allele in B8-C4AQ0-C4B1-DR3. J.�Immunol.�Methods 244, 41–47.

Harteveld, K.L., Losekoot, M., Fodde, R., Giordano, P.C., and Bernini, L.F. (1997). The involvement of Alu repeats in recombination events at the alpha-globin gene cluster: characterization of two alphazero-thalassaemia deletion breakpoints. Hum.�Genet. 99, 528–534.

Herbers, J., Schullerus, D., Chudek, J., Bugert, P., Kanamaru, H., Zeisler, J., Ljungberg, B., Akhtar, M., and Kovacs, G. (1998). Lack of genetic changes at specific genomic sites separates renal oncocytomas from renal cell carcinomas. J.�Pathol. 184, 58–62.

Hess, J.F., Fox, M., Schmid, C., and Shen, C.K. (1983). Molecular evolu-tion of the human adult alpha-globin-like gene region: insertion and deletion of Alu family repeats and non-Alu DNA sequences. Proc.�Natl.�Acad.�Sci.�USA 80, 5970–5974.

Hodgson, G., Hager, J.H., Volik, S., Hariono, S., Wernick, M., Moore, D., Nowak, N., Albertson, D.G., Pinkel, D., Collins, C., Hanahan, D., and Gray, J.W. (2001). Genome scanning with array CGH delineates regional alterations in mouse islet carcinomas. Nat.�Genet. 29, 459–464.

Hogervorst, F.B., Nederlof, P.M., Gille, J.J., McElgunn, C.J., Grippeling, M., Pruntel, R., Regnerus, R., van Welsem, T., van Spaendonk, R., Menko, F.H., Kluijt, I., Dommering, C., Verhoef, S., Schouten, J.P., van’t Veer, L.J., and Pals, G. (2003). Large genomic deletions and duplications in the BRCA1 gene identified by a novel quantitative method. Cancer�Res. 63, 1449–1453.

Hoque, M.O., Lee, J., Begum, S., Yamashita, K., Engles, J.M., Schoenberg, M., Westra, W.H., and Sidransky, D. (2003). High-throughput molecular analysis of urine sediment for the detection of bladder cancer by high-density single-nucleotide polymorphism array. Cancer�Res. 63, 5723–5726.

Hu, X., and Worton, R.G. (1992). Partial gene duplication as a cause of human disease. Hum.�Mutat. 1, 3–12.

Inoue, K., Dewar, K., Katsanis, N., Reiter, L.T., Lander, E.S., Devon, K.L., Wyman, D.W., Lupski, J.R., and Birren, B. (2001). The 1.4-Mb CMT1A duplication/HNPP deletion genomic region reveals unique genome architectural features and provides insights into the recent evolution of new genes. Genome�Res. 11, 1018–1033.

Jagadeeswaran, P., Tuan, D., Forget, B.G., and Weissman, S.M. (1982). A gene deletion ending at the midpoint of a repetitive DNA sequence in one form of hereditary persistence of fetal haemoglobin. Nature 296, 469–470.

Junien, C., Despoisse, S., Turleau, C., Nicolas, H., Picard, F., Le Marec, B., Kaplan, J.C., and de Grouchy, J. (1982). Retinoblastoma, deletion 13q14, and esterase D: application of gene dosage effect to prenatal diagnosis. Cancer�Genet.�Cytogenet. 6, 281–287.

Kajii, T., Ferrier, A., Niikawa, N., Takahara, H., Ohama, K., and Avirachan, S. (1980). Anatomic and chromosomal anomalies in 639 spontaneous abortuses. Hum.�Genet. 55, 87–98.

Kallioniemi, A., Kallioniemi, O.P., Sudar, D., Rutovitz, D., Gray, J.W., Waldman, F., and Pinkel, D. (1992). Comparative genomic hybridiza-tion for molecular cytogenetic analysis of solid tumors. Science 258, 818–821.

Kan, Y.W., Holland, J.P., Dozy, A.M., Charache, S., and Kazazian, H.H. (1975). Deletion of the beta-globin structure gene in hereditary per-sistence of foetal haemoglobin. Nature 258, 162–163.

Kenwrick, S., Patterson, M., Speer, A., Fischbeck, K., and Davies, K. (1987). Molecular analysis of the Duchenne muscular dystrophy region using pulsed field gel electrophoresis. Cell 48, 351–357.

Kim, S.H., Bae, J.H., Chae, J.J., Kim, U.K., Choe, S.J., Namkoong, Y., Kim, H.S., Park, Y.B., and Lee, C.C. (1999). Long-distance PCR-based screening for large rearrangements of the LDL receptor gene in Korean patients with familial hypercholesterolemia. Clin.�Chem. 45, 1424–1430.

Kim, S.W., Lee, K.S., Jin, H.S., Lee, T.M., Koo, S.K., Lee, Y.J., and Jung, S.C. (2003). Rapid detection of duplication/deletion of the PMP22 gene in patients with Charcot-Marie-Tooth disease Type 1A and hereditary neuropathy with liability to pressure palsy by real-time quantitative PCR using SYBR Green I dye. J.�Korean�Med.�Sci. 18, 727–732.

Knight, S.J., Regan, R., Nicod, A., Horsley, S.W., Kearney, L., Homfray, T., Winter, R.M., Bolton, P., and Flint, J. (1999). Subtle chromosomal rearrangements in children with unexplained mental retardation. Lancet 354, 1676–1681.

Knowles, M.A. (1999). Identification of novel bladder tumour suppressor genes. Electrophoresis 20, 269–279.

Koenig, M., Beggs, A.H., Moyer, M., Scherpf, S., Heindrich, K., Bettecken, T., Meng, G., Muller, C.R., Lindlof, M., Kaariainen, H., de la Chapelle, A., Kiuru, A., Savontaus, M.-L., Gilgenkrantz, H., Recan, D., Chelly, J., Kaplan, J.-C., Covone, A.E., Archidiacono, N., Romeo, G., Liechti-Gallati, S., Schneider, V., Braga, S., Moser, H., Darras, B.T., Murphy, P., Francke, U., Chen, J.D., Morgan, G., Denton, M., Greenberg, C.R., van Ommen, G.J., and Kunkel, L.M. (1989). The molecular basis for Duchenne versus Becker muscular dystrophy: correlation of severity with type of deletion. Am.�J.�Hum.�Genet. 45, 498–506.

Laurendeau, I., Bahuau, M., Vodovar, N., Larramendy, C., Olivi, M., Bieche, I., Vidaud, M., and Vidaud, D. (1999). TaqMan PCR-based

Page 11: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

chapter  |  12 Detection of Genomic Duplications and Deletions 181

gene dosage assay for predictive testing in individuals from a cancer family with INK4 locus haploinsufficiency. Clin.�Chem. 45, 982–986.

Lejeune, J. (1960). Mongolism, regressive trisomy. Ann.�Genet. 2, 1–38.Lemmers, R.J., van der Maarel, S.M., van Deutekom, J.C., van der

Wielen, M.J., Deidda, G., Dauwerse, H.G., Hewitt, J., Hofker, M., Bakker, E., Padberg, G.W., and Frants, R.R. (1998). Inter- and intra-chromosomal sub-telomeric rearrangements on 4q35: implications for facioscapulohumeral muscular dystrophy (FSHD) aetiology and diag-nosis. Hum.�Mol.�Genet. 7, 1207–1214.

Li, W.H., and Gojobori, T. (1983). Rapid evolution of goat and sheep globin genes following gene duplication. Mol.�Biol.�Evol. 1, 94–108.

Linn, J.F., Lango, M., Halachmi, S., Schoenberg, M.P., and Sidransky, D. (1997). Microsatellite analysis and telomerase activity in archived tis-sue and urine samples of bladder cancer patients. Int.� J.�Cancer 74, 625–629.

Mansfield, E.S. (1993). Diagnosis of Down syndrome and other aneuploi-dies using quantitative polymerase chain reaction and small tandem repeat polymorphisms. Hum.�Mol.�Genet. 2, 43–50.

Meyer, U.A., and Zanger, U.M. (1997). Molecular mechanisms of genetic polymorphisms of drug metabolism. Annu.�Rev.�Pharmacol.�Toxicol. 37, 269–296.

Mullis, K., Faloona, F., Scharf, S., Saiki, R., Horn, G., and Erlich, H. (1986). Specific enzymatic amplification of DNA in vitro: the polymerase chain reaction. Cold�Spring�Harb.�Symp.�Quant.�Biol. 51, 263–273.

Nagahata, T., Hirano, A., Utada, Y., Tsuchiya, S., Takahashi, K., Tada, T., Makita, M., Kasumi, F., Akiyama, F., Sakamoto, G., Nakamura, Y., and Emi, M. (2002). Correlation of allelic losses and clinicopatholog-ical factors in 504 primary breast cancers. Breast�Cancer 9, 208–215.

Neubauer, A., Neubauer, B., and Liu, E. (1990). Polymerase chain reac-tion based assay to detect allelic loss in human DNA: loss of beta-interferon gene in chronic myelogenous leukemia. Nucleic�Acids�Res. 18, 993–998.

Nupponen, N.N., and Visakorpi, T. (2000). Molecular cytogenetics of prostate cancer. Microsc.�Res.�Tech. 51, 456–463.

Ohno, S., Wolf, U., and Atkin, N.B. (1968). Evolution from fish to mam-mals by gene duplication. Hereditas 59, 169–187.

Orkin, S.H., Alter, B.P., Altay, C., Mahoney, M.J., Lazarus, H., Hobbins, J.C., and Nathan, D.G. (1978). Application of endonuclease map-ping to the analysis and prenatal diagnosis of thalassemias caused by globin-gene deletion. N.�Engl.�J.�Med. 299, 166–172.

Ose, T., and Bush, O.B. Jr. (1962). Erythroblastosis fetalis. Report of cases, with one caused by gene deletion. Bibl.�Haematol. 13, 290–291.

Ottolenghi, S., Lanyon, W.G., Paul, J., Williamson, R., Weatherall, D.J., Clegg, J.B., Pritchard, J., Pootrakul, S., and Boon, W.H. (1974). The severe form of alpha thalassaemia is caused by a haemoglobin gene deletion. Nature 251, 389–392.

Palomares, M., Delicado, A., Lapunzina, P., Ajona, D., Aminosa, C., Arcas, J., Martinez Bermejo, A., Fernandez, L., and Lopez-Pajares, I. (2006). MLPA versus multiprobe FISH comparison of 2 methods for the screening of subtelomere rearrangements in 50 patients with idi-opathic mental retardation. Clin.�Genet. 69, 228–233.

Patau, K., Therman, E., Smith, D.W., and Demars, R.I. (1961). Trisomy for chromosome No. 18 in man. Chromosoma 12, 280–285.

Petrij-Bosch, A., Peelen, T., van Vliet, M., van Eijk, R., Olmer, R., Drusedau, M., Hogervorst, F.B., Hageman, S., Arts, P.J., Ligtenberg, M.J., Meijers-Heijboer, H., Klijn, J.G., Vasen, H.F., Cornelisse, C.J., van ‘t Veer, L.J., Bakker, E., van Ommen, G.J., and Devilee, P. (1997). BRCA1 genomic deletions are major founder mutations in Dutch breast cancer patients. Nat.�Genet. 17, 341–345.

Pinkel, D., Segraves, R., Sudar, D., Clark, S., Poole, I., Kowbel, D., Collins, C., Kuo, W.L., Chen, C., Zhai, Y., Dairkee, S.H., Ljung, B.M., Gray, J.W., and Albertson, D.G. (1998). High resolution analysis of DNA copy number variation using comparative genomic hybridiza-tion to microarrays. Nat.�Genet. 20, 207–211.

Puget, N., Torchard, D., Serova-Sinilnikova, O.M., Lynch, H.T., Feunteun, J., Lenoir, G.M., and Mazoyer, S. (1997). A 1-kb Alu-mediated germ-line deletion removing BRCA1 exon 17. Cancer�Res. 57, 828–831.

Puget, N., Sinilnikova, O.M., Stoppa-Lyonnet, D., Audoynaud, C., Pages, S., Lynch, H.T., Goldgar, D., Lenoir, G.M., and Mazoyer, S. (1999). An Alu-mediated 6-kb duplication in the BRCA1 gene: a new founder mutation? Am.�J.�Hum.�Genet. 64, 300–302.

Rabbitts, T.H., Forster, A., Dunnick, W., and Bentley, D.L. (1980). The role of gene deletion in the immunoglobulin heavy chain switch. Nature 283, 351–356.

Redon, R., Ishikawa, S., Fitch, K.R., Feuk, L., Perry, G.H., Andrews, T.D., Fiegler, H., Shapiro, M.H., Carson, A.R., Chen, W., Cho, E.K., Dallaire, S., Freeman, J.L., Gonzalez, J.R., Gratacos, M., Huang, J., Kalactzopoulos, D., Komua, D., MacDonald, J.R., Maynall, C.R., Mei, R., Montgomery, L., Nishimura, K., Okamura, K., Shen, F., Somerville, M.J., Tchuda, J., Valsessia, A., Woodwark, C., Yang, F., Zhang, J., Zerial, T., Zhang, J., Armengol, L., Conrad, D.F., Estivil, X., Tyler-Smith, C., Carter, N.P., Aburatani, H., Lee, C., Jones, K.W., Scherer, S.W., and Hurles, M.E. (2006). Global variation in copy number in the human genome. Nature 144, 428–429.

Richter, S., Vandezande, K., Chen, N., Zhang, K., Sutherland, J., Anderson, J., Han, L., Panton, R., Branco, P., and Gallie, B. (2003). Sensitive and efficient detection of RB1 gene mutations enhances care for families with retinoblastoma. Am.�J.�Hum.�Genet. 72, 253–269.

Rio, M., Molinari, F., Heuertz, S., Ozilou, C., Gosset, P., Raoul, O., Cormier-Daire, V., Amiel, J., Lyonnet, S., Le Merrer, M., Turleau, C., de Blois, M.C., Prieur, M., Romana, S., Vekemans, M., Munnich, A., and Colleaux, L. (2002). Automated fluorescent genotyping detects 10% of cryptic subtelomeric rearrangements in idiopathic syndromic mental retardation. J.�Med.�Genet. 39, 266–270.

Roelfsema, J.H., White, S.J., Ariyurek, Y., Bartholdi, D., Niedrist, D., Papadia, F., Bacino, C.A., den Dunnen, J.T., van Ommen, G.J., Breuning, M.H., Hennekam, R.C., and Peters, D.J. (2005). Genetic heterogeneity in Rubinstein-Taybi syndrome: mutations in both the CBP and EP300 genes cause disease. Am.�J.�Med.�Genet. 76, 572–580.

Rohlfs, E.M., Puget, N., Graham, M.L., Weber, B.L., Garber, J.E., Skrzynia, C., Halperin, J.L., Lenoir, G.M., Silverman, L.M., and Mazoyer, S. (2000). An Alu-mediated 7.1 kb deletion of BRCA1 exons 8 and 9 in breast and ovarian cancer families that results in alter-native splicing of exon 10. Genes�Chromosomes�Cancer 28, 300–307.

Rooms, L., Reyniers, E., Wuyts, W., Storm, K., van Luijk, R., Scheers, S., Wauters, J., van den Ende, J., Biervliet, M., Eyskens, F., van Goethem, G., Laridon, A., Ceulemans, B., Courtens, W., and Kooy, R.F. (2006). Multiplex ligation-dependent probe amplification to detect subtelom-eric rearrangements in routine diagnosis. Clin.�Genet. 69, 58–64.

Rowland, J.S., Barton, D.E., and Taylor, G.R. (2001). A comparison of methods for gene dosage analysis in HMSN type 1. J.� Med.� Genet. 38, 90–95.

Schouten, J.P., McElgunn, C.J., Waaijer, R., Zwijnenburg, D., Diepvens, F., and Pals, G. (2002). Relative quantification of 40 nucleic acid sequences by multiplex ligation-dependent probe amplification. Nucleic�Acids�Res. 30, e57.

Schughart, K., Kappen, C., and Ruddle, F.H. (1989). Duplication of large genomic regions during the evolution of vertebrate homeobox genes. Proc.�Natl.�Acad.�Sci.�USA 86, 7067–7071.

Page 12: Detection of Genomic Duplications and Deletions - U.N.S.L....Detection of Genomic Duplications and Deletions. Graham R. Taylor. 1. and Carol A. Delaney. 2. 1 Regional Genetics Laboratory

Molecular Diagnostics182

Selzer, R.R., Richmond, T.A., Pofahl, N.J., Green, R.D., Eis, P.S., Nair, P., Brothman, A.R., and Stallings, R.L. (2005). Analysis of chromosome breakpoints in neuroblastoma at sub-kilobase resolution using fine-tiling oligonucleotide array CGH. Genes� Chromosomes� Cancer 44, 305–319.

Shimizu, S., Yatabe, Y., Koshikawa, T., Haruki, N., Hatooka, S., Shinoda, M., Suyama, M., Ogawa, M., Hamajima, N., Ueda, R., Takahashi, T., and Mitsudomi, T. (2000). High frequency of clonally related tumors in cases of multiple synchronous lung cancers as revealed by molecular diagnosis. Clin.�Cancer�Res. 6, 3994–3999.

Shuin, T., Kondo, K., Kaneko, S., Sakai, N., Yao, M., Hosaka, M., Kanno, H., Ito, S., and Yamamoto, I. (1995). Results of mutation analyses of von Hippel-Lindau disease gene in Japanese patients: comparison with results in United States and United Kingdom. Hinyokika� Kiyo 41, 703–707.

Siniscalco, M., Robledo, R., Orru, S., Contu, L., Yadav, P., Ren, Q., Lai, H., and Roe, B. (2000). A plea to search for deletion polymorphism through genome scans in populations. Trends�Genet. 16, 435–437.

Sismani, C., Armour, J.A., Flint, J., Girgalli, C., Regan, R., and Patsalis, P.C. (2001). Screening for subtelomeric chromosome abnormalities in children with idiopathic mental retardation using multiprobe telo-meric FISH and the new MAPH telomeric assay. Eur.�J.�Hum.�Genet. 9, 527–532.

Snijders, A.M., Nowak, N., Segraves, R., Blackwood, S., Brown, N., Conroy, J., Hamilton, G., Hindle, A.K., Huey, B., Kimura, K., Law, S., Myambo, K., Palmer, J., Ylstra, B., Yue, J.P., Gray, J.W., Jain, A.N., Pinkel, D., and Albertson, D.G. (2001). Assembly of microarrays for genome-wide measurement of DNA copy number. Nat.�Genet. 29, 263–264.

Southern, E.M. (1975). Detection of specific sequences among DNA frag-ments separated by gel electrophoresis. J.�Mol.�Biol. 98, 503–517.

Stenson, P.D., Ball, E.V., Mort, M., Phillips, A.D., Shiel, J.A., Thomas, N.S., Abeysinghe, S., Krawczak, M., and Cooper, D.N. (2003). Human Gene Mutation Database (HGMD): 2003 update. Hum.� Mutat. 21, 577–581.

Suminaga, R., Takeshima, Y., Yasuda, K., Shiga, N., Nakamura, H., and Matsuo, M. (2000). Non-homologous recombination between Alu and LINE-1 repeats caused a 430-kb deletion in the dystrophin gene: a novel source of genomic instability. J.�Hum.�Genet. 45, 331–336.

Swensen, J., Hoffman, M., Skolnick, M.H., and Neuhausen, S.L. (1997). Identification of a 14 kb deletion involving the promoter region of BRCA1 in a breast cancer family. Hum.�Mol.�Genet. 6, 1513–1517.

Taylor, C.F., Charlton, R.S., Burn, J., Sheridan, E., and Taylor, G.R. (2003). Genomic deletions in MSH2 or MLH1 are a frequent cause of

T

v

v

V

V

W

W

W

W

Y

hereditary non-polyposis colorectal cancer: identification of novel and recurrent deletions by MLPA. Hum.�Mutat. 22, 428–433.

yagi, S., and Kramer, F.R. (1996). Molecular beacons: probes that fluo-resce upon hybridization. Nat.�Biotechnol. 14, 303–308.

an Essen, A.J., Kneppers, A.L., van der Hout, A.H., Scheffer, H., Ginjaar, I.B., ten Kate, L.P., van Ommen, G.J., Buys, C.H., and Bakker, E. (1997). The clinical and molecular genetic approach to Duchenne and Becker muscular dystrophy: an updated protocol. J.�Med.�Genet. 34, 805–812.

an Loghem, E., Sukernik, R.I., Osipova, L.P., Zegers, B.J., Matsumoto, H., de Lange, G., and Lefranc, G. (1980). Gene deletion and gene duplication within the cluster of human heavy-chain genes. Selective absence of IgG sub-classes. J.�Immunogenet. 7, 285–299.

ortmeyer, A.O., Huang, S.C., Pack, S.D., Koch, C.A., Lubensky, I.A., Oldfield, E.H., and Zhuang, Z. (2002). Somatic point mutation of the wild-type allele detected in tumors of patients with VHL germline deletion. Oncogene 21, 1167–1170.

oskova-Goldman, A., Peier, A., Caskey, C.T., Richards, C.S., and Shaffer, L.G. (1997). DMD-specific FISH probes are diagnostically useful in the detection of female carriers of DMD gene deletions. Neurology 48, 1633–1638.

agner, A., van der Klift, H., Franken, P., Wijnen, J., Breukel, C., Bezrookove, V., Smits, R., Kinarsky, Y., Barrows, A., Franklin, B., Lynch, J., Lynch, H., and Fodde, R. (2002). A 10-Mb paracentric inversion of chromosome arm 2p inactivates MSH2 and is responsible for hereditary nonpolyposis colorectal cancer in a North-American kindred. Genes�Chromosomes�Cancer 35, 49–57.

iencke, J.K., Pemble, S., Ketterer, B., and Kelsey, K.T. (1995). Gene deletion of glutathione S-transferase theta: correlation with induced genetic damage and potential role in endogenous mutagenesis. Cancer�Epidemiol.�Biomarkers�Prev. 4, 253–259.

ijnen, J., van der Klift, H., Vasen, H., Khan, P.M., Menko, F., Tops, C., Meijers Heijboer, H., Lindhout, D., Moller, P., and Fodde, R. (1998). MSH2 genomic deletions are a frequent cause of HNPCC. Nat.�Genet. 20, 326–328.

ilkie, A.O., Higgs, D.R., Rack, K.A., Buckle, V.J., Spurr, N.K., Fischel-Ghodsian, N., Ceccherini, I., Brown, W.R., and Harris, P.C. (1991). Stable length polymorphism of up to 260 kb at the tip of the short arm of human chromosome 16. Cell 64, 595–606.

au, S.C., Bobrow, M., Mathew, C.G., and Abbs, S.J. (1996). Accurate diagnosis of carriers of deletions and duplications in Duchenne/Becker muscular dystrophy by fluorescent dosage analysis. J.� Med.�Genet. 33, 550–558.