Genetic Heterogeneity in Rubinstein-Taybi Syndrome: Mutations in Both the CBP and EP300 Genes Cause Disease (original) (raw)

Am J Hum Genet. 2005 Apr; 76(4): 572–580.

Jeroen H. Roelfsema,1 Stefan J. White,1 Yavuz Ariyürek,1 Deborah Bartholdi,2 Dunja Niedrist,2 Francesco Papadia,3 Carlos A. Bacino,4 Johan T. den Dunnen,1 Gert-Jan B. van Ommen,1 Martijn H. Breuning,1 Raoul C. Hennekam,5 and Dorien J. M. Peters1

Jeroen H. Roelfsema

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Stefan J. White

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Yavuz Ariyürek

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Deborah Bartholdi

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Dunja Niedrist

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Francesco Papadia

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Carlos A. Bacino

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Johan T. den Dunnen

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Gert-Jan B. van Ommen

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Martijn H. Breuning

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Raoul C. Hennekam

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Dorien J. M. Peters

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

1Center for Human and Clinical Genetics, Leiden University Medical Center, Sylvius Laboratory, Leiden, The Netherlands; 2Institute of Medical Genetics, University of Zurich, Zurich; 3Department of Metabolic Diseases and Clinical Genetics, Pediatric Hospital Giovanni XXIII, Bari, Italy; 4Department of Molecular and Human Genetics, Baylor College of Medicine, Houston; and 5Department of Clinical Genetics, Academic Medical Center, Amsterdam

Address for correspondence and reprints: Dr. Dorien J. M. Peters, Center for Human and Clinical Genetics, LUMC, Sylvius Laboratory, Wassenaarseweg 72, 2333AL, Leiden, The Netherlands. E-mail: ln.cmul@sretep.m.j.d

Received 2004 Oct 14; Accepted 2005 Jan 21.

Copyright © 2005 by The American Society of Human Genetics. All rights reserved.

Abstract

CREB-binding protein and p300 function as transcriptional coactivators in the regulation of gene expression through various signal-transduction pathways. Both are potent histone acetyl transferases. A certain level of CREB-binding protein is essential for normal development, since inactivation of one allele causes Rubinstein-Taybi syndrome (RSTS). There is a direct link between loss of acetyl transferase activity and RSTS, which indicates that the disorder is caused by aberrant chromatin regulation. We screened the entire CREB-binding protein gene (CBP) for mutations in patients with RSTS by using methods that find point mutations and larger rearrangements. In 92 patients, we were able to identify a total of 36 mutations in CBP. By using multiple ligation-dependent probe amplification, we found not only several deletions but also the first reported intragenic duplication in a patient with RSTS. We extended the search for mutations to the EP300 gene and showed that mutations in EP300 also cause this disorder. These are the first mutations identified in EP300 for a congenital disorder.

Introduction

Rubinstein-Taybi syndrome (RSTS [MIM 180849]) is a congenital disorder characterized by mental and growth retardation and a wide range of typical dysmorphic features. Facial dysmorphology includes down-slanted palpebral fissures, a broad nasal bridge, a beaked nose, and micrognathia. Particularly noticeable are the broad thumbs and broad big toes. In addition, patients with RSTS have an increased risk of tumor formation. Although various types of tumors have been described, there is an excess of tumors arising from developmental defects and tumors in brain or neural-crest cell–derived tissue (Miller and Rubinstein 1995). Mutations in the gene encoding the CREB-binding protein (CREBBP, also known as “_CBP_”), located on chromosome 16p13.3, were found to be responsible for causing the disorder (Petrij et al. 1995).

The protein, CBP, serves as a transcriptional coactivator (Kwok et al. 1994). It has a transactivation domain but does not specifically bind to DNA. The name of the protein is based on the interaction with the CRE-binding protein (CREB); however, CBP interacts with a large number of other proteins as well. It is thought that CBP acts as an integrator of the signals from various pathways (Goodman and Smolik 2000). Transcription factors downstream from these pathways need to compete with each other for the limited amount of CBP available in the nucleus. The protein forms a physical bridge between the DNA-binding transcription factors and the RNA polymerase II complex. In addition, CBP has intrinsic histone acetyl transferase (HAT) activity (Bannister and Kouzarides 1996). By acetylating histones, CBP opens the chromatin structure at the locus that needs to be expressed, a process essential for gene expression. CBP is also capable of acetylating a large number of other proteins—for example, the transcription factor p53 (Gu and Roeder 1997).

Although RSTS is considered to be an autosomal dominant disorder, patients with RSTS very rarely have children. Almost all mutations, therefore, occur de novo. The mutations found in patients vary from relatively large microdeletions, which remove the entire gene, to point mutations. In addition, five translocations and two inversions disrupting the gene have been reported (Petrij et al. 2000). The microdeletions that remove the entire gene indicate that haploinsufficiency is the ultimate cause of the syndrome. Presumably, at critical moments during development, the amount of CBP drops below a certain threshold because of the loss of one allele. How this loss of the allele actually causes the particular symptoms of RSTS, however, is unclear. Nevertheless, we know from studies of patients with missense mutations and splice-site mutations that affect only the HAT domain of CBP that a loss of HAT activity is sufficient to cause the syndrome (Murata et al. 2001; Kalkhoven et al. 2003).

To elucidate the complete spectrum of mutations, we screened 92 patients with RSTS for point mutations, small deletions or insertions, and large deletions and duplications. Because we could not find mutations in the CBP gene in the majority of our patients, we assumed that the remaining patients have mutations in other genes.

CBP shares homology with another protein, p300, which is encoded by the EP300 gene on chromosome 22q13.2 (Lundblad et al. 1995). Both proteins are particularly homologous at their binding sites for transcription factors, and p300 also has a HAT domain. Like CBP, it serves as a transcriptional coactivator. EP300 is therefore a likely gene candidate to screen, and, indeed, we found three mutations in our analysis of the gene. These are the first mutations found in EP300 for a congenital disorder, and, in addition, they prove that RSTS is a genetically heterogeneous disorder.

Material and Methods

The majority of DNA samples described in this study were sent to us by clinicians in the Netherlands and many other countries in the form of soluble genomic DNA from patients with a clinical diagnosis of RSTS. DNA from the remainder of the patients was isolated from peripheral blood in our laboratory by use of standard protocols. The study was reviewed by the institutional review board of the Leiden University Medical Center. Parents or guardians of patients provided oral or written consent for molecular studies. DNA from 92 patients was analyzed. The 92 patients included some patients who have been presented in previous publications: mutations affecting the HAT domain have been described by Kalkhoven et al. (2003), and an mRNA deletion in patient 127-2 was described by Petrij et al. (2000), although no genomic mutation was found.

Sequence variants were checked using a set of 98 control chromosomes. When possible, DNA from the parents of patients was analyzed to determine whether sequence variants occurred de novo.

Denaturing Gradient Gel Electrophoresis (DGGE)

DGGE was performed with a GC clamp on either the forward or the reverse primer (see table A1 [online only]). To screen the splice sites and branch sites, primers were selected to anneal to the flanking intron sequences and were chosen by use of either WINMELT (Biorad) or MELT-INGENY (Ingeny B.V.) software. All oligonucleotides were synthesized by Sigma-Aldrich. Amplified fragments were analyzed on 9% polyacrylamide gels (acrylamide:bisacrylamide, 37.5:1), with various linear denaturing gradients, optimized for each fragment, on the DCode system from Biorad. Gels were run at 90 V at a constant temperature of 60°C. An acrylamide mixture with 40% formamide and 7 M urea was defined as 100% denaturant, and acrylamide without these denaturing agents was defined as 0% denaturant.

SSCP Analysis

Electrophoresis was performed at room temperature by use of two types of gels. The first type was a polyacrylamide gel (acrylamide:bisacrylamide, 49:1) with 1 × Tris borate EDTA (TBE) without glycerol, and the second type was 0.5 × Mutation Detection Enhancement (National Diagnostics) with 0.6 × TBE and 10% glycerol. During amplification, the fragments for SSCP analysis were radioactively labeled, either by incorporation of α32P-dCTP or by use of primers that were phosphorylated using γ32P-dATP (Amersham). Visualization of the fragments was done using the PhosphorImager (Molecular Dynamics).

Multiple Ligation-Dependent Probe Amplification

Probes were designed for 20 exons of the CBP and EP300 genes. Multiple ligation-dependent probe amplification (MLPA) was performed as described by White et al. (2004); however, for some samples, the amplifications with the multiplex amplifiable probe hybridization (MAPH) and MLPA primers were performed in separate reactions. All samples in which rearrangements were found were tested at least twice.

Sequencing and Restriction Digestions

Sequencing was performed on the ABI 3700 from Applied Biosystems by use of the manufacturer's standard protocol and reagents. Restriction digestions were performed in accordance with the instructions of the manufacturer. Digestions or second sequencing reactions to confirm the first results were done on PCR fragments generated in an independent reaction. The deletion of 8 nt in patient 256-1 was confirmed by PCR with an allele-specific primer, TCCTCCATCTACTAGTAGTG, that skips the deleted part and anneals with two nucleotides after the deletion. The reverse primer has the sequence GTCCTAACCAAATCAAACAG.

Results

Point Mutations and Small Deletions or Insertions in CBP

We screened the entire CBP gene for point mutations and small deletions or insertions, primarily by using DGGE, and screened target sequences that were not suited for DGGE by using SSCP analysis. The complete coding sequence and splice sites of the CBP gene required a total of 49 fragments, of which 40 were screened using DGGE, which covered ∼83% of the coding sequence. Direct sequencing was used to identify the mutation after aberrant bands were found on DGGE or SSCP gels. All mutations were confirmed either by digestion with restriction enzymes, when a restriction-enzyme site was altered, or by a second sequence analysis.

In 92 patients, we found a total of 27 mutations (table 1). The majority is predicted to lead to a premature translation stop, but we also detected five putative missense mutations. Base substitutions leading to a premature stop codon, as well as deletions and insertions leading to frame shifts, can be clearly identified as disease-causing mutations. On the other hand, a change of amino acids is much less clear; however, patients with RSTS, as a rule, have de novo mutations. Since we were able to confirm three of the mutations as de novo, we consider them the most likely to be disease causing. We do not have parental DNA for patients 228-1 and 260-1. All putative missense mutations are at the highly conserved HAT domain of CBP, and the amino acids that are changed have residues that are conserved in both the mouse and the fruit fly (fig. 1).

An external file that holds a picture, illustration, etc. Object name is AJHGv76p572fg1.jpg

Conservation of amino acids that are predicted to change by missense mutations. All five mutations we found that are predicted to change the amino acid residues are situated in the highly conserved HAT domain. The changed residues are conserved in man (Homo sapiens [Hs]), mouse (Mus musculus [Mm]), and the fruit fly (Drosophila melanogaster [Dm]). Numbers indicate individual patients.

Table 1

List of All Mutations Found in the CBP and EP300 Genes

Gene, Type of Mutation,and Individual Exona Mutationa,b Protein Changec
CBP:
Nonsense mutation:
7-1 Exon 2 c.304 C→T Q102X
177-1 Exon 5 c.1237 C→T R413X
212-1d Exon 28 c.4669 C→T Q1558X
27-1 Exon 29 c.4879 A→T K1627X
2-1 Exon 31 c.6010 C→T R2004X
16-1 Exon 31 c.6133 C→T Q2045X
178-3 Exon 31 c.6283 C→T Q2095X
Missense mutation:
209-1d Exon 21 c.3823 G→Ae E1278K
201-1 Exon 26 c.4340 C→Te T1447I
260-1 Exon 26 c.4348 T→C Y1450H
228-1 Exon 27 c.4409 A→G H1470R
2644d Exon 30 c.4991 G→Ae R1664H
Deletion/insertion:
153-1 Exon 2 c.235 del G G79fsX86
199-3 Exon 3 c.904_905 del AGe S302fsX348
205-1 Exon 6 c.1381_1388 del 8e G461fsX469
239-1 Exon 6 c.1481 dup A N494fsX527
203-1 Exon 8 c.1735 dup Ae A581fsX586
57-3 Exon 18 c.3396_3400 del 6 P1132fsX1166
10-1 Exon 18 c.3432_3433 del AG T1144fsX1168
232-1 Exon 21 c.3824 dup T F1275fsX1282
231-1d Exon 25 c.4256_4258 del CT S1419fsX1419
34-3 Exon 27 c.4399 del Ge V1467fsX1467
213-1d Exon 29 c.4837 del Ge V1613fsX1634
Splice-site mutation:
198-3d Exon 20 c.3779 +5 G→Ce
211-3d Exon 22 c.3837 −2 A→Te
47-3 Exon 23 c.3915 −1 G→Ae
39-1d Exon 24 c.4133 +1 G→A
Rearrangementf:
267-1 Del exon 1 c.-198-?_85+? del
36-3 Del exon 1_2 c.-198-?_798+? del
74-1 Del exon 1_19 c.-198-?_3698+? del
15-1 Del exon 1_31 c.-198-?_+1150+? del
41-3 Del exon 1_31 c.-198-?_+1150+? del
127-2g Del exon 2 c.86-?_798+? del
252-1 Del exon 12 c.2159-?_2283+? del
253-1 Del exon 31 c.5173-?_+1150+? del
162-1 Dup exon 1 c.-198-?_85+? dup
EP300:
254-1 Exon 10 c.1942 C→Te R648X
256-1 Exon 15 c.2877_2884 del 8e S959fsX966
149-1 Del exon 1 c.-1200-?_94+? dele

Unless we have an RNA sample from a patient, we cannot check whether a splice-site mutation actually leads to aberrant splicing. These mutations, however, should also comply with the rule that mutations in patients with RSTS occurred de novo. Except for the mutation in patient 39-1, for which parental DNA was not available, we could confirm the mutations in that way. The G→A mutation in the splice-donor site flanking exon 24 in patient 39-1, however, is at the first position, which—in all splice-donor sites, without exception—should be a guanine. The splice-site mutation in patient 211-3 could be analyzed on RNA isolated from a cell line. Subsequent sequence analysis proved that the mutation in the splice-acceptor site flanking exon 22 leads to a deletion of exon 22 in the processed mRNA (Kalkhoven et al. 2003)

Large Deletions and Duplications in CBP

Previous research suggests that ∼10% of the mutations in patients with RSTS are microdeletions affecting the CBP gene (Bartsch et al. 1999; Blough et al. 2000). We performed a FISH analysis using five cosmids spanning the entire gene to detect such deletions, when metaphase chromosome spreads of patients were available (Petrij et al. 2000). The recently developed technique of MLPA can also be used to detect microdeletions in soluble genomic DNA (Schouten et al. 2002). Because that was the type of material available to us for the majority of our patients, we performed MLPA to test the CBP gene.

The resolution of MLPA is related to the number of probes used. We made a set of 20 MLPA probe pairs, covering most of the CBP gene. This allows us to screen for deletions that cannot be detected by FISH. Southern blotting could have been an alternative but is, in our case, impractical—if not impossible—because it requires so much DNA.

The quality of DNA is slightly more critical for MLPA than for a normal PCR; therefore, we could not screen all patients with MLPA who were screened with DGGE and SSCP. In total, we screened 53 patients, and, for controls, we used material from 3 patients with known microdeletions that were detected using FISH, including 1 patient with a deletion of the entire CBP gene. Our MLPA analysis detected those positive controls flawlessly, and we found a number of previously undetected mutations. We found a total of nine new deletions, which ranged from single-exon deletions to the deletion of the entire gene. One deletion, that of exon 2, has been described elsewhere for the RNA level (Petrij et al. 2000). At the time, Southern blots did not reveal a deletion in the genomic DNA; therefore, it was not clear whether this was a genomic deletion or a splicing aberration. This mutation has been found in family 127, which consists of an affected mother and child, one of the very few cases of inherited RSTS.

In addition to the nine deletions, we also found a duplication in one individual. Patient 162-1 has a duplication of the first exon of the CBP gene. How this leads to the inactivation of this allele is not clear, but a disease-causing duplication of the first exon has been described in Opitz syndrome (Winter et al. 2003).

The exon 1 deletions and duplication were confirmed using extra probe pairs—one probe at the promoter region and three probe pairs in intron 1 (see table A2 [online only]).

Mutations in EP300

Point-mutation detection and MLPA analysis of CBP yielded a total of 36 mutations in 92 patients, strongly suggesting that other genes could also be involved in RSTS. The most likely candidate is the EP300 gene, coding for p300, on chromosome 22q13.1. That gene was screened using the same approach. We used 37 DGGE fragments, which covered ∼79% of the coding sequence of EP300, and the remaining part was covered by 10 SSCP fragments. MLPA was performed using a set of 20 exon-specific probe pairs.

Indeed, three inactivating mutations were detected in the EP300 gene (fig. 2). Two mutations were found using DGGE. One mutation, in exon 10, is a transition (c.1942 C→T) that converts the triplet coding for an arginine at position 648 into a stop codon. The other mutation, in exon 15, is a deletion of 8 nt that predicts a frameshift from codon 959, with a stop codon after 7 aa. The exact location of the 8-bp deletion (c.2877_2884) was confirmed with allele-specific PCR. We analyzed DNA from the healthy parents of both patients by use of DGGE and sequencing and confirmed that the mutations occurred de novo. The biological parentage was confirmed by genotyping with 17 independent markers (data not shown). Both mutations lead to predicted proteins less than half their normal size that do not contain the HAT domain. The third mutation, a deletion of the first exon, was found using MLPA. Four probes revealed this deletion—two probes upstream of exon 1, one in exon 1, and the fourth in intron 1—to be close to the first exon. They all showed a decreased signal, whereas a probe in exon 2 showed a normal dosage (fig. 2C). It is probable that this deletion will lead to no expression from the affected allele.

An external file that holds a picture, illustration, etc. Object name is AJHGv76p572fg2.jpg

Mutations in EP300 in patients with RSTS. A, Results for patient 254-1. DGGE analysis of patient 254-1 and the healthy parents shows that only the affected child has the mutation. Subsequent sequence analysis revealed a transition, c.1942 C→T, that predicts the protein change p.Arg648X in this patient. B, Results for patient 256-1. DGGE analysis of family 256 shows a de novo mutation. The allele-specific PCR confirms the exact location of the deletion seen by sequence analysis. The patient has an 8-bp deletion (c.2877_2884 del) with the following sequence (deleted region is in capital letters; nucleotides of allele-specific forward primer are underlined): gcctcctccatctactagtagCACAGAAGtgaat. The forward primer skips the deleted part and anneals with two nucleotides after the deletion. Only DNA from the patient shows a band of 168 bp, whereas, in the lanes for DNA from the healthy parents, only the prominently visible primer dimers can be seen. C, Bar diagram of MLPA results for patient 149-1. MLPA reveals a deletion at the first exon of EP300. The _X_-axis shows the DNA probes. The probes upstream of the first exon are at positions 787–716 and 5–54 bp before the transcription start site (indicated by −787 and −54, respectively). Ex = exon; Int = intron. The _Y_-axis represents the dosage of DNA: a dosage of 1 indicates the presence of the normal amount of DNA—that is, both alleles are present—whereas a dosage of ∼0.5 typically indicates a deletion of one allele. The diagram clearly shows that the deletion runs from the upstream region of exon 1 into intron 1 and that exons 2 and 3 are present for both alleles. The exact size of the deletion is unknown.

Discussion

We undertook a rigorous screening for point mutations, small deletions or insertions, and large deletions and duplications at the coding section of CBP by using genomic DNA from a large set of patients with RSTS. There is neither a predominant type of mutation nor a clear indication of a clustering of mutations within the CBP gene. If we take a look at missense mutations, however, we see that they are all situated in the HAT domain of CBP. We have published some of these mutations elsewhere and have shown that they affect the HAT activity of CBP (Kalkhoven et al. 2003). In addition, two articles each reported a de novo missense mutation in the HAT domain, clearly underpinning the importance of this domain in relation to the disorder (Murata et al. 2001; Bartsch et al. 2002). In contrast to our findings, a study by Coupry et al. (2002) reported four putative missense mutations of which only one was situated in the HAT domain. The sequence variations were not found in the other patients, and the affected residues were conserved in mouse and were therefore considered to be causative for RSTS. Whether these mutations actually arose de novo could not be confirmed.

We have found mutations in less than half of the patients (∼40%), which is comparable to the outcome of the study by Coupry et al. (2002). DGGE and SSCP analyses, together with detection of nucleotide substitutions, are capable of identifying only relatively small deletions and insertions. To detect larger deletions, we chose to perform MLPA for CBP and EP300. We now have shown that MLPA is capable of detecting deletions in the CBP gene that were previously identified by FISH. Because we have probe pairs corresponding to the majority of exons in both CBP and EP300, our MLPA screening also negates the need for Southern blotting. The use of MLPA has increased the power to detect mutations, enabling us to find smaller deletions than could be detected using FISH.

Nevertheless, the combined analysis of our samples with both MLPA and DGGE or SSCP analysis found mutations in less than half of the patients. It is probable that some of the patients we screened will be regarded, upon closer clinical examination, as having a different syndrome that resembles RSTS; however, we think the majority of these patients may be considered as having true RSTS. Diagnosis of the syndrome has been performed by many clinicians, but we did not find that some have a significantly better record, in terms of the number of mutations found, than others. In addition, the set of patients with RSTS who were screened by Coupry et al. (2002) had been carefully ascertained, yet mutations were not found in more than half their patients as well. Either the CBP gene is mutated at parts that we did not screen, such as the promoter or other regulatory elements, or the mutations are in other genes. The unscreened parts of the CBP gene may harbor some mutations, but it is highly unlikely that they contain the majority of RSTS-related mutations that have not yet been identified in these patients. Indeed, RSTS is genetically heterogeneous, since we identified mutations in the EP300 gene.

The finding of the EP300 mutations raises the question of whether there are phenotypic differences associated with EP300 mutations as compared with CBP mutations (fig. 3). The phenotypes of the patients were, in most respects, compatible with classic RSTS: all patients had heavy and arched eyebrows, long eye lashes, a prominent nose with a long hanging columella, and a pouting lower lip. One patient had a small chin. It is interesting that only one patient showed a very mild downward-slanting of the palpebral fissures, and none had the grimacing smile. As mentioned above, all had a shortened and broad thumb and square distal fingertips, and two had visible fetal pads. Also, the big toe was broad in all patients. One patient had a remarkably short first metatarsal bones, giving rise to a very proximal placement of the halluces. Similar dysmorphology has been found in a patient who also has a deleted CBP gene (Petrij et al. 1995). In all three patients with EP300 mutations, the forefoot was broad and the fifth toe was shorter than normal. The small number of patients prevents any of these findings from being firm conclusions.

An external file that holds a picture, illustration, etc. Object name is AJHGv76p572fg3.jpg

Photographs of the three patients with EP300 mutations. A, The face, a hand, and a foot of patient 149-1. B, A hand, a close-up of the fingers with fetal pads, and a foot of patient 254-1. C, A hand, a foot, and an X-ray of the foot of patient 256-1. Patient 256-1 has the overall typical appearance of a patient with RSTS, with the exception of the feet. The feet have abnormally short first metatarsal bones, as can be clearly seen in the X-ray photograph. Although this is not a typical feature, it does appear in some other patients with RSTS, as well as in individuals with mutations in CBP.

Strikingly, the number of identified patients with RSTS who have EP300 mutations, currently three, is small compared with the number of identified patients with RSTS and CBP mutations (36 patients). Possibly, this ratio of 1:12 represents the different chances of mutations occurring in these two genes. Alternatively, the EP300 gene could have a mutation rate equal to that of the CBP gene, with the majority of mutation carriers not yet diagnosed as having RSTS. In view of this latter explanation, it is interesting that we found many more polymorphisms in the EP300 gene, including some that lead to amino acid changes (data not shown). Nevertheless, the majority of point mutations found in the CBP gene are likely to lead to truncated proteins, and two mutations in the EP300 gene are also predicted to truncate the protein, so it is unlikely that the mutation ratio can be entirely explain by differences in the genotype-phenotype relationship. Therefore, we think that the mutation rates of the loci are different.

The CBP gene harbors an instable region around exon 2. This region was designated as instable because all translocation and inversion breakpoints could be found in all patients with RSTS, except one—as could all leukemia breakpoints for which CBP functions as a fusion partner. In addition, this same genomic piece of DNA proved very hard to clone at the time of positional cloning of the RSTS gene (Giles et al. 1997). The deletion of exon 2 and the deletions and duplication of exon 1 may be caused by this instable region. The instability in this region, however, cannot explain the majority of deletions found at the CBP locus, since most of these deletions have their breakpoints elsewhere (Petrij et al. 2000).

Although CBP and p300 have been shown to have overlapping functions, there are subtle but clear differences between the two proteins (Kalkhoven 2004). During embryogenesis, CBP and EP300 have similar but not completely overlapping expression patterns (Partanen et al. 1999). In addition, experiments with F9 teratocarcinoma cell lines showed that retinoic acid signaling is p300 dependent and does not require CBP, whereas cAMP signaling depends on CBP and not p300 (Kawasaki et al. 1998; Ugai et al. 1999). Recent work with transgenic mice indicated the importance of the acetyl transferase function of p300 in myogenesis, but the acetyl transferase function of Cbp does not seem to be necessary for this process (Roth et al. 2003). The skeletal abnormalities found in heterozygous Cbp knockout mice have not been reported for heterozygous Ep300 knockout mice (Tanaka et al. 1997). We, however, do not find very striking phenotypical differences between patients with mutations in the EP300 gene versus the CBP gene.

Double-heterozygous knockout mice for the Cbp and Ep300 genes resemble the homozygous knockout mice for either gene, in that all three types of mice die in utero, a finding which led to the idea that the combined levels of CBP and p300 are critical during development (Yao et al. 1998). Our findings support this hypothesis and reveal that even a relatively small decrease of either protein has significant developmental consequences. However, it is unclear how a decrease of either protein leads to the specific features of RSTS. Perhaps the partial loss of p300 is compensated for by recruitment of CBP, and subsequent depletion of CBP then leads to RSTS. Alternatively, both proteins could be involved in a common function, and, therefore, the total dosage would be required to prevent a syndrome like RSTS. If so, then this common function has a relationship with the HAT activity of the proteins because loss of only the HAT activity of CBP causes RSTS (Murata et al. 2001; Kalkhoven et al. 2003).

Interestingly, there is a direct link between HAT activity and long-term memory. Heterozygous Cbp knockout mice have diminished mental capabilities. Experiments on these knockout mice revealed that inhibiting histone deacetyltransferase could ameliorate the problems with long-term memory seen in the mice (Alarcon et al. 2004). Transgenic mice with a dominant negative Cbp gene, in which only the HAT activity was ablated, also showed problems with long-term memory. Again, this could be reversed by a histone deacetylase inhibitor (Korzus et al. 2004). In view of these data, it is possible that other proteins with HAT activity—or with a function coupled to HAT activity—may also be involved in RSTS. After all, the three mutations we have found in the EP300 gene, together with the CBP gene mutations, still leave us with more than half of the cases of RSTS unaccounted for.

Acknowledgments

We thank the patients and their parents for their cooperation. We thank S. Trompet, for technical assistance, and Dr. P. de Knijff, for genotyping the families. Furthermore, we thank Dr. R. Carrozzo for clinical assistance. This work was funded by grants from the Dutch Cancer Society (RUL2000-2218), the Netherlands Organisation for Scientific Research (NWO) (912-04-047), and the Center of Medical System Biology, established by the Netherlands Genomics Initiative/NWO.

Appendix A

Table A1

Primers used in DGGE or SSCP analysis

Primerb(5′→3′)
Fragmenta Forward Reverse Fragment Gradient or Typec
CBPex1 GC60-TGCCCGGGGCTGTTTTCGCG GACCACGACCCCCGGACG 50–80
CBPex2a GC55-GCTTAGTTTCTCATTTCCAT CTCCTCGTAGAAGCTCCG 20–50
CBPex2b GC60-AACCTTGTTCCAGATGCTGC ACCCAGGCCCTGCTGCAC 40–70
CBPex2c TCTAGTATCAACCCAGGA GCCCTTGTGAAGCCTGATTA SSCP
CBPex2d GCCTCCCAAGCACTGAATCC GC60-TTAAGCTATGGCCAGAGTT 35–65
CBPex2e GC60-CTCCTCAATAGTAACTCT GCCCTGCATGGCTGGAGTAG 35–65
CBPex2f GGGCTGCTGGCAGAGGAAGG GC60-GTAGGAAGTATTGAAAGTGC 40–70
CBPex3 GC55-CTTTTATATTGACTGCGT ATCCACAGACCACAGCAC 30–60
CBPex4 TTATTTTAGTCTCAGATGCA GC60-GGCAAATTCTTCCTGACC 30–60
CBPex5 GC55-TAATTGAAAGTGTGACGATT TCTCATCTCCAAGCTTCTCT 20–50
CBPex6a GC55-TCTCCCTTTTACTTACCTCT CGTCTGGGGCTGGTTCATGT 25–55
CBPex6b CCCCAGCTCCATGCAGCGA GC55-ACTGTTGATTGATTTCTT 30–65
CBPex7 GC55-ACTGTTGATTGATTTCTT GCTCCTGTTGGACTGTCA 25–55
CBPex8 GC55-AGTCACTTCAGTAGATGGAT GC12-CAAACATCTATGAAACTGCA 30–60
CBPex9 GC55-GTAGTTTGATAGAT GTAGAGGCCAGAGCACGGT 30–60
CBPex10a GC55-AGTTTTAACTGTTTCATA ATGCCTTGTTTATGTAAA 10–30
CBPex10b GC55-GAAAACCAAGGAAACAGG GGCATCTTGGGGAACCAG 30–65
CBPex11 GC55-TATATTAAATAACTAGGG CGGGCCCTATGATTCACCACAAACA 20–50
CBPex12 GC55-TTCCTAATGAACCGTTGT ATGTGAGAGGGAGGGCTATC 30–60
CBPex13 GC55-CCGAACTACAGCTCTGGT TCCTAGGGAGCCACGGTC 45–75
CBPex14a GC55-ATTTTACCATACTCTGTC TGCCCGGAAGACGACACA 45–75
CBPex14b GACACCACCTGGGATGACTC GAGTCTTGGCCCAAAAACAG SSCP
CBPex15 GC55-GATCATTTATGTTACCTTGC AGTTGCGATACGCAGTCA 30–65
CBPex16a GC55-ACATAAAGGTGTTTGATA TGACTGAGAGGCTGTGCCGT 10–40
CBPex16b ACATAAAGGTGTTTGATA GC55-CGTAATAAGGTAATGAATA 25–55
CBPex17 GTCACACCAGCAAAGTTATA GTCACACCAGCAAAGTTATA SSCP
CBPex18 GCCAGATGAGACTGGCATTT TGGATTAACCAGGAAAATTCACTT SSCP
CBPex19 GC55-ATAAAAGTATTTCTCCT CGTGCCTTGCCCTAAGACAT 25–55
CBPex20 CTCTGATTGGTGGCTTCGTT GC55-GGCACCGGTACCTTCCTTAT 25–55
CBPex21 GC55-TTGTTTGAAGAACTAGTTACAA CCCACTCCATAAGGAGTAACTT 5–35
CBPex22 GC55-CGCACACACAGACTTCTACA GCCACTGCAACTGCCCCG 35–65
CBPex23 GC55-TGCTGCATTTTGTTGGTTTG CCATGTGTTGAGAGGAACCA 15–45
CBPex24 GC55- CTGTCTCAGCAGTGCTGTTG GCTCGCAGCACTGTAG 30–65
CBPex25 GC55-TGTGCAGAAGCACCTTGT GAGCCCTGGTCTATCCTA 25–55
CBPex26 TGTTGTTTGTTGCTTGTGTTTG CCCATTATTTCACGGAATAAACAT SSCP
CBPex27 CTGGACGTTCATCTGAC GC55-CAAGTATGCGAATGCAAGAA 30–60
CBPex28a GCGTGCATGGCCCTCATC GC55-GCAGTGCTCTCTTCCTTT 20–50
CBPex28b GC55-GAGAGCATTAAGGAACTAG CTCCCAGCCTGCCACCCT 35–65
CBPex29a CCGTGCCAAGTGTGGAGG GC55-CGCGGCTGATGCTGCTTT 35–65
CBPex29b GC55-GTGCCAAGAAGAAGAACAACAAG GGCCAGAGGCACGGCTGC 35–65
CBPex30 GC55-TTGTGTGGGACTAAAGCG CTTCGTCAGACCCCAGGC 50–80
CBPex31a–b GCCTGGTGGTACTGGGTCCC CGATGAGCTGCTTGCACACC SSCP
CBPex31c–d GGCTGCAAACGCAAGACCAA TGCCGAGCCGCTTCCACCGC SSCP
CBPex31e GCCTACCAGCCAGGTGCC GC60-GGCCTGCATGGATATCAC 55–85
CBPex31f–g CCCCAGCAGCAGCCCATG TCGCCATGTTGGGGTTGT SSCP
CBPex31g–h GCGATGGGAGGCCTGAACC CTGCAGGGCTTGCTGGATGTTGG SSCP
CBPex31i GACCCGGAGGCTACCCACC GC60-CCCAATCTGCTGCTTCATCT 45–75
CBPex31j GC60-CGGATTCTGCAGCAACA GAGACGTGGTGTGGCGAAGG 40–70
CBPex31k AGCCCGTCACCACGGATACA GC55-TGAAAGGGAAAAGGTGATGC 40–70
EP300ex1 CCTGGGTGCGGCGCGGGGAC GGAAGAATAAAGGCGCACCG SSCP
EP300ex2a TTGGTTTTGTCATACTTTGA GCTGGCCATGACTTGACCTG SSCP
EP300ex2b TGCTGCGATCTGGTAGTTCC GC55-GTTCATGACTTGATTAGGCA 30–60
EP300ex2c GC55-TGCGGGCATGAATCCTGGAA GC9- ACTCACATGCATGCCGATTG 35–65
EP300ex3 GC55-TCTTAAATTTTATTGCTT GCCAGTTTGAAAACAAGTCT 20–40
EP300ex4 GC17-ATATATTGTTATATCTCT GC60-TGCTAAGTACAGTAACCCTG 40–70
EP300ex5 TCAACAAGTTAGCTATTATT TATAAATGGCACTGGTGAGA SSCP
EP300ex6 CATACTCAGATGTTTCATAA GTTATTAGGATGACACAA SSCP
EP300ex7 GC55-GCTGTTGTATTTATTTCTGT GC7-GTAATGTGACCCAGGGTATA 20–40
EP300ex8 GCTGGCCATGACTTGACCTG CAGGAGATTGAAAACATG SSCP
EP300ex9 GC7-GATATTACAGTGGTAGGA GC60-TGCTAAAGAGGAATAAAA 15–38
EP300ex10a GC60-GTTTATTTTTTCTGTTAC ACTGGAACCATGCCTGCA 15–40
EP300ex10b GC8-GTATAAGATCCAGAAAGAA GC60-CTCTCAAACAGAAATATA 20–48
EP300ex11 GTGGGGTTTGTGTGTGCAGT TGTTCAGGTAGCAAGTATTA SSCP
EP300ex12 GC60-TTCAGATCTAACATTTTGC GC7-ACCAGGTAAAGGCCAAAGA 25–55
EP300ex13 AGGAGATTGAAAACATG AAAAATAAGTGAAAAATCCA SSCP
EP300ex14a GC60-TTCCTTAATTCTGTTCTG GCTCCCCTGAGACCCTGG 15–45
EP300ex14b GC60-GCACAAATGTCTAGTTCTTC GGACTGTGGCCCAGGTGGCA 30–60
EP300ex14c CCAGCAACAACAATTCCA CTGCCTAAATCCAAATCTC SSCP
EP300ex15 ACTCTGCGTGTGTCTCACCT GC60-GTTCCTATACTGAGGTCCTA 20–48
EP300ex16a GC60-ATACTAAAAATTCTTACGTT GGAGATGACTGGGTAGCTGA 15–40
EP300ex16b GC60-CACTGAGTTAAAAACTGA TGACCGCAATTCTGCCCT 30–50
EP300ex17 GAGATGCTTTTAGAGCTTCA TGGAATGTGAATTATCTCTC SSCP
EP300ex18 GC55-ACTTGAGTAATGTTTGATGT GC7-GGTATCCCAGAAAAGTTAAA 20–40
EP300ex19 GC55-CTGTTTCTGACTTGCCATTC TGGCCCATCCGCATGCACTC 30–55
EP300ex20 ACGGAACAGTTCACCCCAGT GC55-GCATAATCACTGGACAACAA 25–50
EP300ex21 GC60-GTATTCTTAAAAAACCTGA AGAAAAGCCAAAGCGTACTG 10–35
EP300ex22 GC60-AAGTTTTCATTTGGTTAAGG GC7-CCAGAGAAAGTAACAACG 25–50
EP300ex23 GTTGTGTAAGCAAAGTTTTG GC55-CAAAATTTACCTTTAAGA 10–32
EP300ex24 GC60-ACCTCAGTAACTTTTAAC GATACCTTGCTTTCATGC 25–50
EP300ex25 TGTTAGAGTAGTTCATGCT GAGCTAGCCACTGTGAGC SSCP
EP300ex26 GC7-TTTTCCTCTTCATTTCTCTT GC60-TAGTTAAAGAAAAAAAGC 10–32
EP300ex27 GC60-TCTTTTCCTTAATGTTCT TCTATTGTCAGCACCTGG 20–50
EP300ex28 GC55-GAAATTCCTATATGTACA TAGAAGTTTCAAAGGAAA 20–40
EP300ex29 GC60-TCCCAAATTACTTAACAA CAACTCTGGGTGGCTGCA 15–45
EP300ex30 GC55-AAAAAAAGAGACTGTCTG CCCTGCGGAGCTCACCTC 30–60
EP300ex31a GC60-TGAATGACTTAAATCTTG GCTGCAGCCTGCTGGTTG 20–45
EP300ex31b GC60-ATGACCACAAAATGGAGA GGCAGTGCTTGGCATGGT 36–66
EP300ex31c GC60-CAAGGGTTGCAAACGGAAAA GGGGTGGTTGGCTGTTGGCC 40–70
EP300ex31d CACTCCTGCCACTCCAAC GC60-TCTGGTGTTGGATTGGCCTT 45–75
EP300ex31e AGCAGCGGAGACGCAGCG GC60-CTGATACCTACCTGGCCCA 40–65
EP300ex31f GC60-TCAACCTTTGAACATGGCTC CCCCCTGCTGACCTGGCATG 35–65
EP300ex31g ACAACCCATCCCTGGGCAGC GC60-CATCATTTGCTGTCGTCT 40–70
EP300ex31h GC60-GATGAACATGAACCACAACA CAGGGGACCCCATCTGTTGC 32–62
EP300ex31i GC60-AACAAGGAAATATGGGACAG GGGGAAACGTGGTGTGGAGA 35–65
EP300ex31j CCTTCCCCAAGGATGCAG GC60-TATGTCTAGTGTACTCTG 35–65
EP300ex31k CCTGGGACTCAGCACCGA GC60-AATAATTTTTTTGCTCCCAA 20–45

Table A2

Sequences of the Probes Used in the MLPA[Note]

Sequence Location
GGCAAGTTTGGAGGGGCAGATAACTAACT CBP promoter
GGGATGGGAACACCTCGGGTATTCGTA CBP promoter
GCAGGTGAAAATGGCTGAGAACTTGCT CBP exon 1
GGACGGACCGCCCAACCCCAAAAGAGCCAAA CBP exon 1
TAAAGTGCAGCGTGTGGGTTTGTTTGGTAATGTCCTACGC CBP intron 1
CTAGTGGCACTCCAGGAGGGAAGAAG CBP intron 1
GCCAAAGGTAGGGCAGTGTGTTTTCTGGAT CBP intron 1a
TGGCGACACATTGACCAGGCTGCATTTCT CBP intron 1a
GACGTTGTCGTACCTTGTAAGTCCT CBP intron 1b
AGGACTTACTGGGAGTGCTAGTCTGATTCT CBP intron 1b
CCTGATGAGCTGATACCCAATGGAGGAGAAT CBP exon 2
TAGGCCTTTTAAACAGTGGGAACCTTGTTCCAG CBP exon 2
CCCAGTTAGCCAGCAAACAGAGCAT CBP exon 3
GGTCAACAGTTTGCCCACCTTCCCTACAGAT CBP exon 3
GAATTGTACCCACACAAGCAATTGCAACAG CBP exon 4
GCCCCACTGCAGATCCTGAAA CBP exon 4
GTGCATCTTCACGACAAATCATCTCTCATT CBP exon 5
GGAAGAACTGCACACGACATGACTGTCCTG CBP exon 5
CCAGCTAGTGGAATTCAAAACACAATTGGTTCTGTTGGCACA CBP exon 6
GGGCAACAGAATGCCACTTCTTTAAGTAACCC CBP exon 6
GAACGATGGCTCCAACTCTGGTAACATT CBP exon 8
GGAACCCTCAGCACTATACCAACAGCAGCTCCTC CBP exon 8
CAACACCTGATCCCGCAGCTCT CBP exon 9
AAAGGATCGCCGCATGGAAAACCTG CBP exon 9
GGATGAATTCATTTAACCCCATGTCCT CBP exon 12
TGGGGAACGTCCAGTTGCCACAAGCACCCAT CBP exon 12
CGGCTGCTGGCATGCCATCTCTCCA CBP exon 14
GCACACGACACCACCTGGGATGAC CBP exon 14
GGACCTGACGTACCTGTGCTGGAAATGAAG CBP exon 15
ACGGAGACCCAAGCAGAGGACACTGAGCCCGATCCT CBP exon 15
CAAGTTAAAGAAGAAACAGACATAGCAGAGCAGAAATCAG CBP exon 16
AACCAATGGAAGTGGATGAAAAGAAACCTGAAGTGAAAGT CBP exon 16
GCTTGCAGAGGTCTTTGAGCAGGAAATTGACCCT CBP exon 17
GTCATGCAGTCCCTTGGATATTGCTG CBP exon 17
GAATCCCATGGACCTCTCCACCATCAAGCGGAAGCT CBP exon 18
GGACACAGGGCAATACCAAGAGCCCTGGCAGTAC CBP exon 18
CACAGACTTTGTGCTGCTATGGGAAGCAGCT CBP exon 19
GTGTACCATTCCTCGCGATGCTGCCTACTACAGCTAT CBP exon 19
GTATCATTTCTGTGAGAAGTGTTTCACAG CBP exon 20
AGATCCAGGGCGAGAATGTGACCCTGGGTGACGACCCTT CBP exon 20
GTTGATTGCAAGGAGTGTGGCCGGAAGATGCATCAGAT CBP exon 22
TTGCGTTCTGCACTATGACATCATTTGGCCTTCAGG CBP exon 22
GTTTGTGGATTCTGGGGAAATGTCTGAATCTT CBP exon 25
TCCCATATCGAACCAAAGCTCTGTTTGCTTTTGAGGAA CBP exon 25
CAAGCAACTGAAGACAGGCTCACCAGT CBP exon 28
GCCAAGGAACTGCCCTATTTTGAAGGT CBP exon 28
CAGCAGCGGATTCTGCAGCAACA CBP 3′ UTR
GCAGATGAAGCAGCAGATTGGG CBP 3′ UTR
CATGAACACCCGCAACGTGCCTCA CBP 3′ UTR
GCAGAGTCTGCCTTCTCCTACCTCAGCA CBP 3′ UTR
GAATTTGGGTCTGGTGCCGAAACATCTTTCCACTGCT EP300 promoter-787
GTGATTTTGTTTTCTGGTCAGAGTAATAATGCTCG EP300 promoter-787
CTACCGCTAGCCTGTAAGGAGGAT EP300 promoter-54
TCGGCAGAGGGAAGAAACAACAGCCG EP300 promoter-54
GCCGAAGAAGAGATTTCCTGAGGATTCT EP300 exon 1
GGTTTTCCTCGCTTGTATCTCCG EP300 exon 1
TTGCAGAATATTTTCTGTTGAGGAATAGGGTG EP300 intron 1
CTTTCCACTCTTCGAAATTTTCTCTCCT EP300 intron 1
GGGACTAACCAATGGTGGTGATATTAATCAGCT EP300 exon 2
TCAGACAAGTCTTGGCATGGTACAAGATG EP300 exon 2
CCATATACTCAGAATCCTGGACAGCAGATTGGAGCCA EP300 exon 3
GTGGCCTTGGTCTCCAGATTCAGACAAA EP300 exon 3
GCTGATCCAGAGAAGCGCAAGCTCATCCA EP300 exon 4
GCAGCAGCTTGTTCTCCTTTTGCATGCTC EP300 exon 4
CCAACCTAAGCACTGTTAGTCAGATTGATCCCAGCTCCAT EP300 exon 6
AGAAAGAGCCTATGCAGCTCTTGGACTACCCTATCA EP300 exon 6
GTGCTAGTCCTATGGGAGTAAATGGAGGTGTAGG EP300 exon 7
AGTTCAAACGCCGAGTCTTCTTTCTGACTCAATGTTGCAT EP300 exon 7
CAGCAGCTCAACCATCCACTACT EP300 exon 8
GGAATTCGGAAACAGTGGCACGA EP300 exon 8
CGAAGGACCAGACTACAGAAGCAGAACAT EP300 exon 10
GCTACCAAATGCTGCAGGCATGGTTCC EP300 exon 10
GCCCTCTACCTGACCCAAGTATGAT EP300 exon 11
CCGTGGCAGTGTGCCAAACCAGATGATG EP300 exon 11
CAGCCTTCCAACCAGGGCCAGTTCCT EP300 exon 13
TCCTCAGACTCAGTTCCCATCACAGG EP300 exon 13
CAGGTATCAAATCCTCCATCTACTAGTAGCACA EP300 exon 15
GAAGTGAATTCTCAGGCCATTGCTGAG EP300 exon 15
CACTGATGCCAACTTTGGAGGCACTTTACC EP300 exon 17
GTCAGGATCCAGAATCCCTTCCCTTTCGTCAACCTGT EP300 exon 17
CATGGATCTTTCTACCATTAAGAGGAAGTTAGACACT EP300 exon 18
GGACAGTATCAGGAGCCCTGGCAGTATGTCG EP300 exon 18
GAAGTGTTTCAATGAGATCCAAGGGGAG EP300 exon 20
AGCGTTTCTTTGGGGGATGACCCTTCC EP300 exon 20
GTTGAATGTACAGAGTGCGGAAGAAAGAT EP300 exon 22
GCATCAGATCTGTGTCCTTCACCATGAGATCAT EP300 exon 22
GGTTGCCATCTACCAGACTTGGCACCTTTCTAGAG EP300 exon 23
AATCGTGTGAATGACTTTCTGAGGCGACAGAATC EP300 exon 23
CATATCTTACCTCGATAGTGTTCATTTCTTCCGTCCTAAAT EP300 exon 26
GCTTGAGGACTGCAGTCTATCATGAAATCCTAATTGGA EP300 exon 26
CATATTTGGGCATGTCCACCAAGTGAGGGAGATGATTATATCT EP300 exon 27
TCCATTGCCATCCTCCTGACCAGAAGATACCCAAGCCC EP300 exon 27
GAGCAGCCTGAGTAGGGGCAACAAGAA EP300 exon 29
GAAACCCGGGATGCCCAATGTATCTAA EP300 exon 29
GGTATCAGCCCACTCAAACCAGGCACTGT EP300 exon 31
GTCTCAACAAGCCTTACAAAACCTTTTGCG EP300 exon 31

Electronic-Database Information

Accession numbers and URLs for data presented herein are as follows:

References

Alarcon JM, Malleret G, Touzani K, Vronskaya S, Ishii S, Kandel ER, Barco A (2004) Chromatin acetylation, memory, and LTP are impaired in CBP+/− mice: a model for the cognitive deficit in Rubinstein-Taybi syndrome and its amelioration. Neuron 42:947–959 [PubMed] [Google Scholar]

Bannister AJ, Kouzarides T (1996) The CBP coactivator is a histone acetyltransferase. Nature 384:641–643 [PubMed] [Google Scholar]

Bartsch O, Locher K, Meinecke P, Kress W, Seemanova E, Wagner A, Ostermann K, Rodel G (2002) Molecular studies in 10 cases of Rubinstein-Taybi syndrome, including a mild variant showing a missense mutation in codon 1175 of CREBBP. J Med Genet 39:496–501 [PMC free article] [PubMed] [Google Scholar]

Bartsch O, Wagner A, Hinkel GK, Krebs P, Stumm M, Schmalenberger B, Bohm S, Balci S, Majewski F (1999) FISH studies in 45 patients with Rubinstein-Taybi syndrome: deletions associated with polysplenia, hypoplastic left heart and death in infancy. Eur J Hum Genet 7:748–756 [PubMed] [Google Scholar]

Blough RI, Petrij F, Dauwerse JG, Milatovich-Cherry A, Weiss L, Saal HM, Rubinstein JH (2000) Variation in microdeletions of the cyclic AMP-responsive element-binding protein gene at chromosome band 16p13.3 in the Rubinstein-Taybi syndrome. Am J Med Genet 90:29–34 [PubMed] [Google Scholar]

Coupry I, Roudaut C, Stef M, Delrue MA, Marche M, Burgelin I, Taine L, Cruaud C, Lacombe D, Arveiler B (2002) Molecular analysis of the CBP gene in 60 patients with Rubinstein-Taybi syndrome. J Med Genet 39:415–421 [PMC free article] [PubMed] [Google Scholar]

den Dunnen JT, Antonarakis SE (2001) Nomenclature for the description of human sequence variations. Hum Genet 109:121–124 [PubMed] [Google Scholar]

Giles RH, Petrij F, Dauwerse HG, den Hollander AI, Lushnikova T, van Ommen GJ, Goodman RH, Deaven LL, Doggett NA, Peters DJ, Breuning MH (1997) Construction of a 1.2-Mb contig surrounding, and molecular analysis of, the human CREB-binding protein (CBP/CREBBP) gene on chromosome 16p13.3. Genomics 42:96–114 [PubMed] [Google Scholar]

Goodman RH, Smolik S (2000) CBP/p300 in cell growth, transformation, and development. Genes Dev 14:1553–1577 [PubMed] [Google Scholar]

Gu W, Roeder RG (1997) Activation of p53 sequence-specific DNA binding by acetylation of the p53 C-terminal domain. Cell 90:595–606 [PubMed] [Google Scholar]

Kalkhoven E (2004) CBP and p300: HATs for different occasions. Biochem Pharmacol 68:1145–1155 [PubMed] [Google Scholar]

Kalkhoven E, Roelfsema JH, Teunissen H, den Boer A, Ariyurek Y, Zantema A, Breuning MH, Hennekam RC, Peters DJ (2003) Loss of CBP acetyltransferase activity by PHD finger mutations in Rubinstein-Taybi syndrome. Hum Mol Genet 12:441–450 [PubMed] [Google Scholar]

Kawasaki H, Eckner R, Yao TP, Taira K, Chiu R, Livingston DM, Yokoyama KK (1998) Distinct roles of the co-activators p300 and CBP in retinoic-acid-induced F9-cell differentiation. Nature 393:284–289 [PubMed] [Google Scholar]

Korzus E, Rosenfeld MG, Mayford M (2004) CBP histone acetyltransferase activity is a critical component of memory consolidation. Neuron 42:961–972 [PMC free article] [PubMed] [Google Scholar]

Kwok RP, Lundblad JR, Chrivia JC, Richards JP, Bachinger HP, Brennan RG, Roberts SG, Green MR, Goodman RH (1994) Nuclear protein CBP is a coactivator for the transcription factor CREB. Nature 370:223–226 [PubMed] [Google Scholar]

Lundblad JR, Kwok RP, Laurance ME, Harter ML, Goodman RH (1995) Adenoviral E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP. Nature 374:85–88 [PubMed] [Google Scholar]

Miller RW, Rubinstein JH (1995) Tumors in Rubinstein-Taybi syndrome. Am J Med Genet 56:112–115 [PubMed] [Google Scholar]

Murata T, Kurokawa R, Krones A, Tatsumi K, Ishii M, Taki T, Masuno M, Ohashi H, Yanagisawa M, Rosenfeld MG, Glass CK, Hayashi Y (2001) Defect of histone acetyltransferase activity of the nuclear transcriptional coactivator CBP in Rubinstein-Taybi syndrome. Hum Mol Genet 10:1071–1076 [PubMed] [Google Scholar]

Partanen A, Motoyama J, Hui CC (1999) Developmentally regulated expression of the transcriptional cofactors/histone acetyltransferases CBP and p300 during mouse embryogenesis. Int J Dev Biol 43:487–494 [PubMed] [Google Scholar]

Petrij F, Dauwerse HG, Blough RI, Giles RH, van der Smagt JJ, Wallerstein R, Maaswinkel-Mooy PD, van Karnebeek CD, van Ommen GJ, van Haeringen A, Rubinstein JH, Saal HM, Hennekam RC, Peters DJ, Breuning MH (2000) Diagnostic analysis of the Rubinstein-Taybi syndrome: five cosmids should be used for microdeletion detection and low number of protein truncating mutations. J Med Genet 37:168–176 [PMC free article] [PubMed] [Google Scholar]

Petrij F, Giles RH, Dauwerse HG, Saris JJ, Hennekam RC, Masuno M, Tommerup N, van Ommen GJ, Goodman RH, Peters DJ, Breuning MH (1995) Rubinstein-Taybi syndrome caused by mutations in the transcriptional co-activator CBP. Nature 376:348–351 [PubMed] [Google Scholar]

Roth JF, Shikama N, Henzen C, Desbaillets I, Lutz W, Marino S, Wittwer J, Schorle H, Gassmann M, Eckner R (2003) Differential role of p300 and CBP acetyltransferase during myogenesis: p300 acts upstream of MyoD and Myf5. Embo J 22:5186–5196 [PMC free article] [PubMed] [Google Scholar]

Schouten JP, McElgunn CJ, Waaijer R, Zwijnenburg D, Diepvens F, Pals G (2002) Relative quantification of 40 nucleic acid sequences by multiplex ligation-dependent probe amplification. Nucleic Acids Res 30:e57 [PMC free article] [PubMed] [Google Scholar]

Tanaka Y, Naruse I, Maekawa T, Masuya H, Shiroishi T, Ishii S (1997) Abnormal skeletal patterning in embryos lacking a single Cbp allele: a partial similarity with Rubinstein-Taybi syndrome. Proc Natl Acad Sci USA 94:10215–10220 [PMC free article] [PubMed] [Google Scholar]

Ugai H, Uchida K, Kawasaki H, Yokoyama KK (1999) The coactivators p300 and CBP have different functions during the differentiation of F9 cells. J Mol Med 77:481–494 [PubMed] [Google Scholar]

White SJ, Vink GR, Kriek M, Wuyts W, Schouten J, Bakker B, Breuning MH, den Dunnen JT (2004) Two-color multiplex ligation-dependent probe amplification: detecting genomic rearrangements in hereditary multiple exostoses. Hum Mutat 24:86–92 [PubMed] [Google Scholar]

Winter J, Lehmann T, Suckow V, Kijas Z, Kulozik A, Kalscheuer V, Hamel B, Devriendt K, Opitz J, Lenzner S, Ropers HH, Schweiger S (2003) Duplication of the MID1 first exon in a patient with Opitz G/BBB syndrome. Hum Genet 112:249–254 [PubMed] [Google Scholar]

Yao TP, Oh SP, Fuchs M, Zhou ND, Ch’ng LE, Newsome D, Bronson RT, Li E, Livingston DM, Eckner R (1998) Gene dosage-dependent embryonic development and proliferation defects in mice lacking the transcriptional integrator p300. Cell 93:361–372 [PubMed] [Google Scholar]


Articles from American Journal of Human Genetics are provided here courtesy of American Society of Human Genetics