Top Banner
Recent insights into genotype–phenotype relationships in patients with Rett syndrome using a fine grain scale Rosa Angela Fabio a, *, Barbara Colombo b , Silvia Russo c , Francesca Cogliati c , Maura Masciadri c , Silvia Foglia b , Alessandro Antonietti b , Daniela Tavian b,d a Department of Cognitive Science, Education and Cultural Studies, University of Messina, via Concezione 8, 98122 Messina, Italy b Department of Psychology, Catholic University of the Sacred Heart, Largo Gemelli 1, 20123 Milano, Italy d Laboratory of Cellular Biochemistry and Molecular Biology-CRIBENS, Catholic University of the Sacred Heart, Piazza Buonarroti 30, 20145 Milano, Italy c Cytogenetics and Molecular Genetics Laboratory, Istituto Auxologico Italiano, via Ariosto 13, 20145 Milano, Italy 1. Introduction Rett syndrome (RTT; OMIM 312750) is a developmental disorder, almost exclusively affecting females, resulting in severe mental retardation and neurological disability. RTT progresses from the latent onset of symptoms to a more complete expression of the disorder. In late infancy, after a period of superficially normal but subtly flawed development, RTT patients undergo striking developmental regression. RTT is characterized by the loss of pre-existing hand use – such as object reach, grasp, and manipulation, and by the appearance of distinctive hand stereotypies – such as hand wringing, tapping, and Research in Developmental Disabilities 35 (2014) 2976–2986 ARTICLE INFO Article history: Received 15 April 2014 Received in revised form 9 July 2014 Accepted 14 July 2014 Available online Keywords: Rett syndrome MECP2 Cognitive processes Emotion Motor behaviour Autonomy ABSTRACT Mutations in MECP2 gene cause Rett syndrome (RTT), a neurodevelopmental disorder affecting around 1 in 10,000 female births. The clinical picture of RTT appears quite heterogeneous for each single feature. Mutations in MECP2 gene have been associated with the onset of RTT. The most known gene function consists of transcriptional repression of specific target genes, mainly by the binding of its methyl binding domain (MBD) to methylated CpG nucleotides and recruiting co-repressors and histone deacetylase binding to DNA by its transcription repressor domain (TRD). This study aimed at evaluating a cohort of 114 Rett syndrome (RTT) patients with a detailed scale measuring the different kinds of impairments produced by the syndrome. The sample included relatively large subsets of the most frequent mutations, so that genotype–phenotype correlations could be tested. Results revealed that frequent missense mutations showed a specific profile in different areas of impairment. The R306C mutation, considered as producing mild impairment, was associated to a moderate phenotype in which behavioural characteristics were mainly affected. A notable difference emerged by comparing mutations truncating the protein before and after the nuclear localization signal; such a difference concerned prevalently the motor-functional and autonomy skills of the patients, affecting the management of everyday activities. ß 2014 Published by Elsevier Ltd. * Corresponding author. E-mail address: [email protected] (R.A. Fabio). Contents lists available at ScienceDirect Research in Developmental Disabilities http://dx.doi.org/10.1016/j.ridd.2014.07.031 0891-4222/ß 2014 Published by Elsevier Ltd.
11

Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

Mar 12, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

Research in Developmental Disabilities 35 (2014) 2976–2986

Contents lists available at ScienceDirect

Research in Developmental Disabilities

Recent insights into genotype–phenotype relationships in

patients with Rett syndrome using a fine grain scale

Rosa Angela Fabio a,*, Barbara Colombo b, Silvia Russo c, Francesca Cogliati c,Maura Masciadri c, Silvia Foglia b, Alessandro Antonietti b, Daniela Tavian b,d

a Department of Cognitive Science, Education and Cultural Studies, University of Messina, via Concezione 8, 98122 Messina, Italyb Department of Psychology, Catholic University of the Sacred Heart, Largo Gemelli 1, 20123 Milano, Italyd Laboratory of Cellular Biochemistry and Molecular Biology-CRIBENS, Catholic University of the Sacred Heart, Piazza Buonarroti 30, 20145

Milano, Italyc Cytogenetics and Molecular Genetics Laboratory, Istituto Auxologico Italiano, via Ariosto 13, 20145 Milano, Italy

A R T I C L E I N F O

Article history:

Received 15 April 2014

Received in revised form 9 July 2014

Accepted 14 July 2014

Available online

Keywords:

Rett syndrome

MECP2

Cognitive processes

Emotion

Motor behaviour

Autonomy

A B S T R A C T

Mutations in MECP2 gene cause Rett syndrome (RTT), a neurodevelopmental disorder

affecting around 1 in 10,000 female births. The clinical picture of RTT appears quite

heterogeneous for each single feature. Mutations in MECP2 gene have been associated with

the onset of RTT. The most known gene function consists of transcriptional repression of

specific target genes, mainly by the binding of its methyl binding domain (MBD) to

methylated CpG nucleotides and recruiting co-repressors and histone deacetylase binding

to DNA by its transcription repressor domain (TRD). This study aimed at evaluating a

cohort of 114 Rett syndrome (RTT) patients with a detailed scale measuring the different

kinds of impairments produced by the syndrome. The sample included relatively large

subsets of the most frequent mutations, so that genotype–phenotype correlations could be

tested. Results revealed that frequent missense mutations showed a specific profile in

different areas of impairment. The R306C mutation, considered as producing mild

impairment, was associated to a moderate phenotype in which behavioural characteristics

were mainly affected. A notable difference emerged by comparing mutations truncating

the protein before and after the nuclear localization signal; such a difference concerned

prevalently the motor-functional and autonomy skills of the patients, affecting the

management of everyday activities.

� 2014 Published by Elsevier Ltd.

1. Introduction

Rett syndrome (RTT; OMIM 312750) is a developmental disorder, almost exclusively affecting females, resulting in severemental retardation and neurological disability. RTT progresses from the latent onset of symptoms to a more completeexpression of the disorder. In late infancy, after a period of superficially normal but subtly flawed development, RTT patientsundergo striking developmental regression. RTT is characterized by the loss of pre-existing hand use – such as object reach,grasp, and manipulation, and by the appearance of distinctive hand stereotypies – such as hand wringing, tapping, and

* Corresponding author.

E-mail address: [email protected] (R.A. Fabio).

http://dx.doi.org/10.1016/j.ridd.2014.07.031

0891-4222/� 2014 Published by Elsevier Ltd.

Page 2: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–2986 2977

mouthing (Fabio, Giannatiempo, Antonietti, & Budden, 2009b). Post-regression patients, even though persons with severeintellectual disabilities, often regain social interest (Antonietti, Castelli, Fabio, & Marchetti, 2008; Castelli, Antonietti, Fabio,Lucchini, & Marchetti, 2013; Fabio, Castelli, Marchetti, & Antonietti, 2013) and are relatively stable for an extended period.After this period, progressive motor deterioration occurs in the form of weakness, wasting, and dystonia. During this period,crude self-feeding capabilities may be retained, but voluntary hand use is generally exceedingly limited and handstereotypies become pervasive (Smeets, Pelc, & Dan, 2011; Weng, Bailey, & Cobb, 2011). Some studies showed that patientsoften remain visually attentive to objects and people, tracking their movements and even showing preferences by ‘‘eyepointing’’ (Fabio, Antonietti, Marchetti, & Castelli, 2009a). There is a cohort of RTT patients, who does not fit into the above‘‘classical’’ form and have been grouped in according to an atypical RTT phenotype, which is characterized by the age of theregression onset (frusta forma, late regression onset, neonatal encephalopathy) or by for the presence of verbal speech(Preserved Speech Variant). Renieri et al. (2009) proposed the term ‘‘Zappella variant’’ rather than ‘‘preserved speechvariant’’ to described milder forms of RTT, because other aspects, besides speech, are involved.

Considering this range of behavioural patterns, it is important to analyze if specific phenotypic symptoms in RTT arerelated to specific genotypes. Amir et al. (1999) firstly demonstrated that mutations in the MECP2 gene, mapping to Xq28region, are associated with RTT. MECP2 encodes Methyl-CpG-binding 2 protein, which belongs to a large family of DNA-binding protein characterized by the presence of methyl binding domain (MBD) that selectively binds 5-methylcytosineresidues in symmetrically positioned CpG dinucleotides (Fan & Hutnick, 2005; Nikitina et al., 2007). Additionally, atranscriptional repression domain (TRD), which interacts with various co-repressor complexes (Jones et al., 1998; Nan,Campoy, & Bird, 1997; Nan et al., 1998), a bipartite nuclear localization signal (NLS) and a WW domain binding region (WDR)(Buschdorf & Stratling, 2004) have been recognized in the protein structure. As regards the functional role of these domains,it has been showed (Jones et al., 1998) that after binding to methylated CpG nucleotides, MECP2 can recruit, by its TRD,histone deacetylases to a transcriptional repressor complex, silencing target genes; the WDR seems to be involved inprotein–protein interaction (Buschdorf and Stratling, 2004). Two alternatively spliced MECP2 transcripts, with a differentATG, have been characterized (Mnatzakanian et al., 2004) determining the production of MeCP2E1 and MeCP2E2 proteinsisoforms. To date, it has been estimated that 80% of RTT patients carry mutations within MECP2 gene, up to 95% consideringthe classical form. Recently CDKL5 and FOXg1 gene mutations have been identified in girls affected by atypical RS with early-onset seizures and by some congenital variant forms (Ariani et al., 2008; Scala et al., 2005).

More than 200 different mutations of MECP2 gene have been reported in the Rett Base (IRSA MECP2 Variation Database;http://mecp2.chw.edu.au/mecp2/), but eight mutations (Arg106Trp, Arg133Cys, Thr158Met, Arg168X, Arg255X, Arg270X,Arg294X, Arg306Cys) affect around 67% of RTT females. A remaining 10% of RTT cases show a large group of C-terminalframeshift mutations.

Several studies have reported genotype-phenotype correlations, but with conflicting results. Most authors reporting datafrom different cohorts of RTT patients (Auranen et al., 2001; Chae, Hwang, & Kim, 2002; Weaving et al., 2003) demonstratedthat no correlation exists between missense vs. truncating mutations, whereas others (Colvin et al., 2004; Huppke, Held,Hanefeld, Engel, & Laccone, 2002; Monros et al., 2001) reported that the truncating defects are more severe than themissense ones. Studies aimed at comparing mutations affecting the different functional domains share the opinion thatdefects affecting the C-terminal domain give milder clinical score (Colvin et al., 2004; Huppke et al., 2002). Differences inclustering the mutations, the heterogeneity in size of the analyzed cohorts, the selected clinical parameters, and variation inthe age of subjects are likely to explain the conflicting results.

Another important factor modulating phenotype expression is the randomization of X-inactivation; when present, it isexpected to influence phenotypical severity. Nevertheless, most studies reported that RTT patients’ lymphocytes fail to showa skewed X chromosome inactivation (XCI) (Amir et al., 2000; Hoffbuhr, Moses, Jerdonek, Naidu, & Hoffman, 2002; Van denVeyver & Zoghbi, 2001). Yet, Knudsen (Knudsen et al., 2006) found a significant increment of skewed XCI in RTT patients andtheir mothers. However, the inclusion of this datum in genotype–phenotype correlation studies remains controversialbecause of the bias of tissue mosaicism. In fact, the brain cannot respect the same ratio as those found in blood or fibroblasts,which are the investigated tissues. Moreover a role of SNPs and CNVs as modifier elements of phenotype in patients carryingthe same MECP2 mutations should be considered (Artuso et al., 2011).

Given that conflicting findings about genotype–phenotype relationships in RTT are still under discussion (Halbach et al.,2011), it seems worthwhile to further investigate this relationship in order to overcome some methodological flaws(Halbach et al., 2011; Ham, Kumar, Deeter, & Schanen, 2005). In this study, we investigated the effect of MECP2 mutations onthe phenotypic variability within a group of 114 RTT patients, focusing on specific methodological issues. More precisely, ourstudy was performed taking into account what was recommended by Ham et al. (2005) concerning the weak points of theprevious studies. Firstly, biases coming from missing data, multiple testing, and age differences were minimized. Secondly, amain new feature was added to this analytic examination of the RTT phenotype, by using the Rett Assessment Rating Scale(R.A.R.S.) (Vignoli et al., 2010), a specific instrument devised to measure the intensity of RTT symptoms and to provide aspecific RTT behavioural profile. R.A.R.S. is a standardized scale, which provides sub-scores concerning the specificsymptomatic areas and an overall score that expresses the severity of the clinical condition of the RTT patient. Both the sub-scores and the overall score are normalized measures.

The specific aim of this study was to correlate disease-causing genetic mutations to the variety and intensity of thedetailed symptomatic parameters assessed by R.A.R.S. and to provide insights into the effect of MECP2 mutations on RTTphenotype.

Page 3: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–29862978

2. Methods

2.1. Patients

One hundred and fourteen Italian RTT patients participated in the study. The RTT patients’ ages ranged from 2 to 36 years(mean = 14.10 years, SD = 9.75). To analyze possible influences of the age on the phenotype–genotype relationships,participants were divided into three groups corresponding, respectively, to the first, second and third tertiles of thedistribution of the ages. The first group included patients ranging from 2 to 7 years, the second from 8 to 14, and the thirdfrom to 15 to 36.

Patients were involved in the study thanks to the collaboration of the Italian Rett Association (A.I.R.), who contactedpatients’ families and asked them to being involved in the study; 72% of the contacted families gave their consensus andprovided researchers with the requested data.

We reconstructed the stories of the patients by examining the clinical documentation of each patient; such anexamination led to conclude that, until the regression phase of RTT, all participants showed a typical psycho-physicaldevelopment proven by the routine paediatric check-ups. When evident RTT symptoms emerged, patients were evaluatedby medical specialists operating in hospitals or RTT centres. The onset of regression occurred at a mean age of 20 months(range = 12–38 months, median = 18 months). Parents suspected a developmental problem at a mean age of 12 months(range = 5–20 months, median = 15 months).

At the time of the study, all patients were living at home in various districts of Italy. The youngest participants wereattending nursery schools or primary schools, whereas the oldest were involved in activities carried out in socio-educationalstructures. All the patients were in the post-regression phase of the disorder; 92% of them were persons with severeintellectual handicap and were unable to use speech. All patients showed little or no purposeful hand use and pervasive handstereotypies. The ability to walk was preserved in 52% of the patients.

According to the paperwork provided by the parents, all the patients examined for MECP2 mutations were positive.The reports of the genetic examination showed that DNA was extracted from lymphocytes of whole blood and thatmutational screening was carried out by direct sequencing or by DHPLC screening followed by the sequencing ofheteroduplexes.

2.2. Instrument

The structure of R.A.R.S. is similar to that of C.A.R.S (The Childhood Autism Rating Scale) (Schopler, Reichler, &Rochen-Renner, 1998), G.A.R.S. (Gilliam Autism Rating Scale) (Gilliam, 1995), and A.S.D.S. (Asperger Syndrome Diagnostic

Scale) (Myles & Simpson, 1998), well-known instruments devised to assess the presence/absence of symptomscharacterizing pervasive developmental disorders included in the same nosographic category of RTT (AmericanPsychiatric Association, 2000).

The construction of the items of the scale was carried out following the diagnostic criteria for RTT proposed by DSM-IV-TR (American Psychiatric Association, 2000) and following recent research and clinical experience. A total of 30items, representative of the profile of RTT, were devised. Table 1 shows the categories used as a basis for the specificsubscales.

Each item concerns a specific phenotypic characteristic. It reports the description of four increasing levels of severity ofthe issue in question. Each item comes with a brief glossary that explains its meaning in a few words. Each item has to berated on a 7-point scale as follows: 1 = within normal limits; 2 = infrequent or low abnormality; 3 = frequent or medium-highabnormality; 4 = strong abnormality. Intermediate rates can be endorsed (for example, when the answer is between thepoint 2 and the point 3, the 2.5 point has to be marked). Table 2 reports an example of how R.A.R.S. works.

R.A.R.S. differs from other existing scale since it includes a high number of items – whereas, for instance, the Pineda, Kerr,and Percy scales, employed by Bebbington et al. (2008), are composed, respectively, only by 10, 19, 15 items and the ClinicalSeverity Score used by Neul et al. (2008) by 13 items. Secondly, it allows evaluators to discriminate the level of severity of theclinical manifestations in a fine grain manner, since each manifestation has to be evaluated on a 7-point scale (in comparison,for instance, to the 5-point scale employed in the Clinical Severity Score or to even less points in the other above mentionedscales).

For each item, the evaluator has to circle the number that corresponds to the point that best describes the RTT patient.After a patient has been rated on each of the 30 items, a total score is computed by summing the individual ratings. This totalscore allows the evaluator to identify the level of severity of RTT, conceptualized as a continuum, ranging from mildsymptoms to severe deficits (Erlandson & Hagberg, 2005). Scores for each subscale are computed as well, by summing up thescores of each item belonging to the corresponding area (sensory, cognitive, and so on).

The R.A.R.S. was obtained by a standardization procedure, involving a sample of 220 RTT patients, proving that theinstrument is statistically valid and reliable (Vignoli et al., 2010). More precisely, normal distribution analyses of the scoreswere computed and the mean scores of the scale were similar to the median and to the mode. Skewness and kurtosis values,calculated for the distribution of the total score, were .110 and .352, respectively. The distribution was found to be normal.Cronbach’s a was used to determine the internal consistency for the whole scale and subscales. Total a was .912, and theinternal consistency of the subscales was high (a varying from .811 to .934).

Page 4: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

Table 1

Subscales of the R.A.R.S.

SUBSCALE TOPICS

Sensory Visual perception

Auditory perception

Cognitive Attention

Spatial orientation

Temporal orientation

Memory

Eye contact

Verbal skills

Gestures

Motor-structural and functional Foot problems

Scoliosis

Walking

Hand use

Emotional Understanding

Expression

Autonomy Toileting

Dressing

Eating and drinking

Physical characteristics Breathing irregularities

Seizures

Bruxism

Abnormal ocular movements

Epilepsy

Aerophagy

Muscle Abnormality

Feeding Problems

Behavioural characteristics Mood variability

Hyperactivity

Anxiety

Aggression

Table 2

An example of the items of the RA.R.S.

1 She normally uses nonverbal communication, which is appropriate considering her age and situation.

1.5

2 She uses a relatively mildly abnormal nonverbal communication. She may only point vaguely, or reach for what she wants,

in the same situations where a child of the same age may point or gesture more specifically to indicate what he or she wants.

2.5

3 She uses a moderately abnormal nonverbal communication. She is generally unable to express needs or desires nonverbally,

and cannot understand the nonverbal communication of others.

3.5

4 She shows a severely abnormal use of nonverbal communication and shows no awareness of the meanings associated with

the gestures or facial expressions of others.

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–2986 2979

2.3. Procedure

The whole examination of the clinical documentation was carried out, for each participant, by the same clinician, who hada long and wide experience in diagnosing and assessing RTT patients. The same clinician scored the R.A.R.S. for each patientbefore examining the clinical and genetic documentation, so to be blind of the mutation of the patient.

3. Results

All the mutations identified in our sample are reported in Fig. 1, which schematically represents the MECP2 gene with its 4exons and functional domains. There were 38 different mutations in all, including private and hot-spot mutations, groupedin 10 missense, 3 large deletions, 2 in frame deletions, whereas the remaining 23 were truncating. It is worth noting that inour sample 3 out of the 4 most frequent mutations (T158M, R306C, R133C) were included in the most frequent variationsidentified in a large database (Percy et al., 2007). Patients carrying hot-spot sequence alterations (Arg106Trp, Arg133Cys,Thr158Met, Arg168X, Arg255X, Arg270X, Arg294X and Arg306Cys) represent 65% of the whole cohort, and 14% of mutationsaffects the C-terminal (10 deletions, 2 nonsense, 2 in-frame deletions and 2 missense). Two large deletions include exons 3and 4, whereas a third deletion lacks exon 3.

Page 5: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

[(Fig._1)TD$FIG]

Fig. 1. Scheme of MECP2 gene with its 4 exons and functional domains: NTS, MBD, ID, TRD and CTS. MECP2 gene mutations identified in Rett patients are

reported. (For interpretation of the references to color in the text citation of this figure, the reader is referred to the web version of this article.)

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–29862980

3.1. Frequent missense mutations

3.1.1. Mutation R133C

Patients with R133C mutation present a mild to moderate level of impairment and can be considered less severe thanpatients showing other common missense variations (Table 3 and Fig. 2). Motor, autonomy and behavioural areas appearedto be the ones where the differences were more pronounced (Fig. 3). In Fig. 4a, it is possible to observe the specific R.A.R.Sprofile of R133C patients compared with the scores of patients with mutations associated to mild, moderate, and severedeficits; cognitive, physical and sensory areas were affected the most.

Table 3

Evaluation of common and rare missense mutations by bio-informatic prediction tools and by known scales devised to assess the level of severity of RTT

Common

missense

mutations

SIFT PolyPhen-2 Pmut Homology Score ISS Score CSS Score R.A.R.S.

R133C Damaging Probably damaging Pathol. Conserved 13.7 (10 PTs) 18.1 (12 PTs) 58.428 (7 PTs)

P152R Tolerated Probably damaging Pathol. Conserved 25.2 (5 PTs) ND 65.357 (7 PTs)

T158M Damaging Probably damaging Neutral Conserved 21 (15 PTs) 22.3 (30 PTs) 62.5 (17 PTs)

R306C Damaging Probably damaging Pathol. Conserved 15 (12 PTs) 21.6 (21 PTs) 65.35 (7 PTs)

Rare missense

mutations

SIFT PolyPhen-2 Pmut Homology Score ISS Score CSS Score R.A.R.S.

R106W Damaging Probably damaging Pathol. Conserved ND (4 PTs) 24.8 (9 PTs) 63.5 (2 PTs)

N126Y Damaging Probably damaging Pathol. Conserved ND ND 63 (1 PT)

P251L Damaging Probably damaging Pathol. Conserved ND ND 68 (1 PT)

P302L Damaging Probably damaging Pathol. Conserved ND ND 62 (1 PT)

P376S Tolerated Benign Neutral Not conserved ND ND 80 (1 PT)

SIFT, score from 1 to 0;�0.05 is damaging; PolyPhen, score from 0 to 1;�0.8 is probably damaging; Pmut, gene variations can be considered as pathogical or

neutral; Homology, a multiple sequence alignment of mammalian MECP2 proteins was used as input for ClustalW; Score ISS; reference: Halbach et al

(2011); Score CSS; reference: Neul et al. (2008); PTs, Patients

.

.

Page 6: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

[(Fig._3)TD$FIG]

Fig. 3. Differences between frequent genetic variants in the R.A.R.S. subscales.

[(Fig._2)TD$FIG]

Fig. 2. R.A.R.S total mean scores in frequent missense mutations as compared with those of mild, moderate and severe mutations.

[(Fig._4)TD$FIG]

Fig. 4. Mean scores on the R.A.R.S. subscales in frequent missense mutations as compared with those of mild, moderate and severe mutations.

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–2986 2981

Page 7: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–29862982

3.1.2. Mutation P152R

The effects of this mutation were moderate (Table 3 and Fig. 2). The cognitive area appeared to be the most damagedcompared to other genetic variants (Fig. 4b), whereas the motor area was the least damage.

3.1.3. Mutation T158M

This mutation is quite frequent in RTT population. Data from our sample confirmed that it can be considered a moderatemutation (Table 3 and Fig. 2). The cognitive and emotional areas were the least compromised, with values close to those of amild impairment (Fig. 4c). Colvin et al. (2004) reported a high variability among these patients. So, we used P–P plots on oursamples to explore the distribution of the scores both of the R.A.R.S overall scale and of the subscales. R.A.R.S. Scores dopresent a positive skew, mainly due to sensory, motor, autonomy and physical characteristics subscales (Fig. 1S).

Supplementary Fig. 1S related to this article can be found, in the online version, at http://dx.doi.org/10.1016/j.ridd.2014.07.031.

3.1.4. Mutation R306C

In our sample, R306C mutation corresponded to a medium-high level of impairment, which allows us to classify themutation as having moderate effects (Table 3 and Fig. 2). As shown in Fig. 4d, behavioural characteristics were affected themost, with a score that was more close to severe than moderate damages.

From the comparison of R133C and R306C mutations, which have been described in literature as mild but with a differenteffect on language skills, emerged that the R306 mutation affects most considered areas, but only the difference in the motorarea was statistically significant (t12 =�2.44; p< .05) (Fig. 5).

3.2. Rare missense mutations

In our sample it was possible to identify 5 rare genetic variants (Table 3). A particularly interesting result concerns thehigh score of the patient who showed the P376S mutation. A comparison of the R.A.R.S. subscales between this specificpatient and benign mutation is reported in Fig. 6.

3.3. Truncating mutations

A comparison was made within four different clusters of truncating mutations to test the hypothesis that they affectfunctional domains differently: clustering 1: truncating within TRD vs. truncating after TRD; clustering 2: truncating within[(Fig._5)TD$FIG]

Fig. 5. Differences in mean scores on the R.A.R.S. subscales between R133C and R306C mutations.[(Fig._6)TD$FIG]

Fig. 6. Comparison between R.A.R.S. subscales mean scores of patient with P376S mutation with the mean scores for benign mutations.

Page 8: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–2986 2983

NLS vs. truncating after NLS; clustering 3: truncating within MBD vs. truncating after MBD and clustering 4: truncatingwithin MBD vs. truncating after MBD and within NLS.

As shown in Table 4, patients with a truncating mutation after NLS manifested a lower degree of impairment in themotor-functional and in the autonomy subscales than patients with a truncating mutation within NLS (clustering 2). Thesame was true for the total score of R.A.R.S. Sixty-nine patients were included in the analysis. Group A consisted of 45patients with a truncating mutation within NLS (black, green and blue mutations with the exclusion of Arg294X in Fig. 1) andgroup B comprised 24 patients with a truncating mutation after NLS mutation (Arg294X and red mutations in Fig. 1). The twoclusters were similar with respect to age of the patients (crosstabulation between age levels and clusters: x2 (2,N = 67) = 1.46; p = .481).

In clustering 1, 3 and 4 no significant difference emerged in R.A.R.S. scores (Tables 1S–3S).Supplementary Tables 1S–3S related to this article can be found, in the online version, at http://dx.doi.org/10.1016/

j.ridd.2014.07.031.

4. Discussion

RTT is a monogenic disorder mainly due to de novo mutations in the MECP2 gene and manifesting as a large variety ofphenotypes showing from very severe to mild impairments. In this study, disease-causing mutations have been related tothe variety and intensity of the detailed symptomatic parameters assessed by R.A.R.S. so to provide insights into the effect ofMECP2 mutations on RTT phenotype. The study showed that R.A.R.S. can help in identifying possible specific aspects of thephenotype of the RTT, which are related to genotype.

In our cohort of patients, most of missense mutations (R133C, P152R and T158M) were clustered in the DNA-binding domain of MeCP2. This is a very important region for association with methylated DNA. In our study, patientscarrying R133C mutation showed mild damages, whereas patients with P152R and T158M mutations showedmoderate damages. Our results are in line with those reported by numerous authors, referring to the first variationdescribed (R133C) (Bebbington et al., 2008; Colvin et al., 2004; Halbach et al., 2011; Neul et al., 2008; Percy et al.,2007).

Moreover, our data showed that R133C patients present a score similar to moderate mutations, referring to thecognitive area. Recently, Mellen, Ayata, Dewell, Kriaucionis, and Heintz (2012) identified MeCP2 as the majorhydroxymethylcytosine (5hmC) binding protein in the brain and showed that the R133C mutation preferentiallyimpairs 5hmC binding. It was also demonstrated that the genomic distribution of 5hmC is cell specific (Mellen et al.,2012). It is well known that patients carrying the R133C mutation are characterized by delayed onset regression,with improved speech and motor skills (Bebbington et al., 2008), whereas for other characteristics – asbreathing abnormalities, sleep problems, mood disturbances, and epilepsy – no significant difference is evidentwith patients bearing other missense mutations (Bebbington et al., 2008). The observation that the distribution of5hmC is different depending on cell type, and that the R133C disease causing mutation can impact preferentially5hmC binding, could explain the specific phenotype associated with altered MeCP2 function in patients carrying thisspecific mutation.

Considering patients with P152R mutation, no data have been reported previously, with the exception of Halbach et al.(2011)Halbach et al.’s study. Unfortunately, bio-informatic prediction tools are not useful in determining the pathogeniceffect of P152R mutation. Indeed, it is considered pathological by Pmut, probably damaging (score 0.999) by Poliphen-2 andtolerated by SIFT (Table 3). Our results showed that patients harbouring P152R mutation were moderately affected and thatthe cognitive area was the most damaged, whereas the motor area the least.

On the contrary, the T158M variant is very common in RTT and in our cohort (17 patients) as well. Colvin et al. (2004)reported a high variability among these patients but he failed to identify potential drivers of variability. Using P–P plots

Table 4

Scores in the different subscales of the R.A.R.S. according to clustering 2.

Subscale Truncating within NLS Truncating after NLS t p

M SD M SD

Sensory 0.29 1.04 �0.32 0.87 2.50 .01

Cognitive 0.17 0.99 �0.26 0.96 1.85 .07

Motor 0.40 0.93 �0.50 0.91 4.02 < .001

Emotion 0.34 1.19 �0.31 0.81 2.42 .02

Autonomy 0.31 0.71 �0.50 1.12 3.71 <.001

Physical Features �0.06 0.9 0.18 1.33 �0.89 .37

Behavioural Features 0.182 0.99 �0.09 1.15 1.06 .29

Total score 71.10 11.61 61.98 12.49 3.02 <.005

Clustering 2: sixty-nine patients. Group A consisted of 45 patients with a truncating mutation within NLS (black, green and blue mutations with the

exclusion of Arg294X in Fig. 1) and group B comprised 24 patients with a truncating mutation after NLS mutation (Arg294X and red mutations in Fig. 1)

.
Page 9: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–29862984

analysis we were able to point out the specific areas that are affected by this variability (sensory, motor autonomy andphysical characteristics subscales), and also to highlight that the cognitive and the emotional areas were the leastdamaged.

Our data on frequent missense mutations located in methyl CpG DNA binding domain clearly showed that the first twovariations (R133C and P152R) affect mainly the cognitive functions (as attention, memory, spatial and temporal orientation),whereas the last (pT158M) is associated with a minor damage in the same area.

The R306C mutation, previously found to confer milder features, resulted to be associated to a more severephenotype in our cohort of RTT patients. This data confirmed what reported by Neul et al. (2008) and contrast Halbachet al. (2011)Halbach’s results. In particular, behavioural characteristics were affected the most, with a mean scoreresembling that of severe damage. The R306C mutation is localized in the C-terminal extremity of the transcriptionalrepression domain (TRD). Very recently it has been demonstrated that MeCP2 R306C mutation abolished theinteraction between MeCP2 and the NCoR/SMRT co-repressor complexes (Lyst et al., 2013). This interaction isrequired for histone deacetylation-mediated gene repression. Mice carrying this mutation exhibited an RTT-likephenotype that resembled that of MeCP2-null mice. It is possible that residual activity of MeCP2 R306 protein mightmitigate the severity of this mutation in humans. Nevertheless, functional studies will need to be performed in orderto understand the molecular interaction between MeCP2 and NCoR/SMRT in human brain (Banerjee, Romero-Lorenzo,& Sur, 2013).

Also considering other sets of mutations, our study, thanks to the fine grain analysis of the phenotype characteristicsallowed by R.A.R.S., succeeded in highlighting a specific profile for each genotype variant. Considering rare MeCP2 missensemutations, the analysis carried out thanks to R.A.R.S. revealed that they all can be considered as producing moderateimpairments, with the exception of the patient harbouring the P376S mutation. These results are consistent with commonbio-informatic prediction tools (Pmut, Poliphen-2 and SIFT) for all rare mutations, with the exception of P376S variations.According to these software, the P376S mutation should not be considered as pathogenic, but patient’s clinical data andanalyses based on R.A.R.S. revealed a severe RTT phenotype. Conforti et al. (2003) hypothesized that P376S mutation couldyield severe deficits and should not be considered as a simple polymorphism although it is localized in the WDR domain,since it has never been detected in control subjects.

Comparisons between the truncating mutations differently affecting functional domains induce to support the ideathat the crucial factor that leads to different phenotypes is the integrity of NLS. We suppose that if the protein canpenetrate into the nucleus and link to Methylated CpG, it maintains a residual role causing a milder clinical damage. Therelevant different degree of affection observed in the two groups of truncating within NLS vs. truncating after NLSmutations in Clustering 2 can also be assigned to the role of MECP2 in condensing chromatin by methylation independentDNA binding. The protein region involved in this role goes from the end of MBD to the N-terminal 294 residue, whereas theC-terminal region is not required (Erlandson & Hagberg, 2005; Lyst et al., 2013; Mellen et al., 2012). The differentialphenotypic effects produced by the two kinds of mutations were clearly shown by the difference in the total R.A.R.S.scores. Furthermore, it is worth noting that the effects produced by this specific kind of mutations are particularly strongin specific areas. Even though in 6 subscales out of 7, as well as in the total scores of the R.A.R.S., patients with a truncatingmutation within NLS obtained higher scores than patients with a truncating mutation after NLS, only in the motor-functional and in the autonomy subscales were the differences statistically significant. Thus, there was evidence thatthese mutations affect selectively the management of everyday activities. Clustering 3 and 4 failed to provide interestingresults, but they supported the notion that in the cases where MBD is not damaged the effects appear similar to thosegiven by more precocious mutations.

Possible limitations to the study have to be considered. First, even though it has been recently reported that clinicalseverity in RTT is independent of X-inactivation status (Neul et al., 2008), we have to admit that the inactivation of X was nottaken into account. Moreover, another putative modifier effect modulating mutation effect could be the presence of a denovo CNV or SNP (Artuso et al., 2011). Second, some analyzed clusters included samples with a limited number of patients.

The most important innovation introduced in the study was the use of a set of fine grain measures to assess the level ofseverity of the RTT symptoms. Most of previous studies assessed the phenotypic features of RTT patients mainly through fewindices, global evaluation, or qualitative descriptions, which are associated to extensive subjectivity. By contrast, ourattempt was to apply an instrument, which allowed us to assess a complete and systematic range of the symptomaticmanifestations of RTT, in order to have a detailed view of each kind of symptoms, expressed through quantitativestandardized measures and recorded according to well-defined operationalized descriptions. Since the same instrument wasapplied to the whole sample following the same procedure, a well-structured and homogeneous set of data recorded in alarge sample of patients could be analyzed.

5. Conclusion

Our investigation into genotype–phenotype relationships of MECP2 mutations in RRT using R.A.R.S. provided moreinsight into disease specific profiles associated with common missense and truncating mutations. The picture that emergedis coherent, allowing us to motivate the puzzling findings reported in the literature, and leads us to conclude that a specifickind of genotypes can be associated to the levels of the symptoms showed by RTT patients.

Page 10: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–2986 2985

Acknowledgments

The collaboration of the Italian association of parents of RTT patients ‘‘A.I.R.’’ (Associazione Italiana Rett) and the supportfrom the ‘‘EuroRETT E-RARE network’’ are gratefully acknowledged.

References

American Psychiatric Association (2000). Diagnostic and statistical manual of mental disorders (DSM-IV-TR). Washington, DC: American Psychiatric Association.Amir, R. E., Van den Veyver, I. B., Wan, M., Tran, C. Q., Francke, U., & Zoghbi, H. Y. (1999). Rett syndrome is caused by mutations in X-linked MeCP2, encoding

methyl-CpG-binding protein 2. Nature Genetics, 23, 185–188.Amir, R. E., Van den Veyver, I. B., Schultz, R., Malicki, D. M., Tran, C. Q., Dahle, E. J., et al. (2000). Influence of mutation type and X chromosome inactivation on Rett

syndrome phenotypes. Annals of Neurology, 47, 670–679.Antonietti, A., Castelli, I., Fabio, R. A., & Marchetti, A. (2008). Understanding emotions and mental states from faces and pictures in Rett syndrome. In M. Balconi

(Ed.), Emotional face comprehension. Neuropsychological perspectives (pp. 205–232). Hauppauge, NY: Nova Science Publishers.Ariani, F., Hayek, G., Rondinella, E., Artuso, R., Mencarelli, M. A., Spanhol-Rosseto, A., et al. (2008). FOXG1 is responsible for the congenital variant of Rett syndrome.

American Journal of Human Genetics, 83, 89–93.Artuso, R., Papa, F. T., Grillo, E., Mucciolo, M., Yasui, D. H., Dunaway, K. W., et al. (2011). Investigation of modifier genes within copy number variations in Rett

syndrome. Journal of Human Genetics, 57, 342–344.Auranen, M., Vanhala, R., Vosman, M., Levander, M., Varilo, T., Hietala, M., et al. (2001). MECP2 gene analysis in classical Rett syndrome and in patients with Rett-

like features. Neurology, 56, 611–617.Banerjee, A., Romero-Lorenzo, E., & Sur, M. (2013). MeCP2: Making sense of missense in Rett syndrome. Cell Research, 23, 1244–1246.Bebbington, A., Anderson, A., Ravine, D., Fyfe, S., Pineda, M., de Klerk, N., et al. (2008). Investigating genotype–phenotype relationships in Rett syndrome using an

international data set. Neurology, 70, 868–875.Buschdorf, J. P., & Stratling, W. H. (2004). A WW domain binding region in methyl-CpG-binding protein MeCP2: Impact on Rett syndrome. Journal of Molecular

Medicine, 82, 135–143.Castelli, I., Antonietti, A., Fabio, R. A., Lucchini, B., & Marchetti, A. (2013). Do Rett syndrome persons possess Theory of Mind? Some evidence from not-treated girls.

Life Span and Disability, 16, 157–168.Chae, J. H., Hwang, Y. S., & Kim, K. J. (2002). Mutation analysis of MECP2 and clinical characterization in Korean patients with Rett syndrome. Journal of Child

Neurology, 17, 33–36.Colvin, L., Leonard, H., De Klerk, N., Davis, M., Weaving, L., Williamson, S., et al. (2004). Refining the phenotype of common mutations in Rett syndrome. Journal of

Medical Genetics, 41, 25–30.Conforti, F. L., Mazzei, R., Magariello, A., Patitucci, A., Gabriele, A. L., Muglia, M., et al. (2003). Mutation analysis of the MECP2 gene in patients with Rett syndrome.

American Journal of Medical Genetics – Part A, 117, 184–187.Erlandson, A., & Hagberg, B. (2005). MECP2 abnormality phenotypes: Clinicopathologic area with broad variability. Journal of Child Neurology, 20, 727–732.Fabio, R. A., Antonietti, A., Marchetti, A., & Castelli, I. (2009). Attention and communication in Rett syndrome. Research in Autism Spectrum Disorders, 3, 329–335.Fabio, R. A., Giannatiempo, S., Antonietti, A., & Budden, S. (2009). The role of stereotypies in overselectivity process in Rett syndrome. Research in Developmental

Disabilities, 30, 136–145.Fabio, R. A., Castelli, I., Marchetti, A., & Antonietti, A. (2013). Training communication abilities in Rett syndrome through reading and writing. Frontiers in

Psychology, 4(911), 1–9. http://dx.doi.org/10.3389/fpsyg.2013.00911Fan, G., & Hutnick, L. (2005). Methyl-CpG binding proteins in the nervous system. Cell Research, 15, 255–261.Gilliam, J. E. (1995). Gilliam autism rating scale. Austin, TX: PRO-ED.Halbach, N. S. J., Smeets, E. E. J., van den Braak, N., van Roozendaal, K. E., Blok, R. M., Schrander-Stumpel, C. T., et al. (2011). Genotype–Phenotype relationships as

prognosticators in Rett syndrome should be handled with care in clinical practice. American Journal of Medical Genetics – Part A, 158A, 340–350.Ham, A. L., Kumar, A., Deeter, R., & Schanen, N. C. (2005). Does genotype predict phenotype in Rett syndrome? Journal of Child Neurology, 20, 768–778.Hoffbuhr, K. C., Moses, L. M., Jerdonek, M. A., Naidu, S., & Hoffman, E. P. (2002). Associations between MeCP2 mutations, X-chromosome inactivation, and

phenotype. Mental Retardation and Developmental Disabilities Research Reviews, 8, 99–105.Huppke, P., Held, M., Hanefeld, F., Engel, W., & Laccone, F. (2002). Influence of mutation type and location on phenotype in 123 patients with Rett syndrome.

Neuropediatrics, 33, 105–108.Jones, P. L., Veenstra, G. J., Wade, P. A., Vermaak, D., Kass, S. U., Landsberger, N., et al. (1998). Methylated DNA and Mecp2 recruit histone deacetylase to repress

transcription. Nature Genetics, 19, 187–191.Knudsen, G. P., Neilson, T. C., Pedersen, J., Kerr, A., Schwartz, M., Hulten, M., et al. (2006). Increased skewing of X chromosome inactivation in Rett syndrome

patients and their mothers. European Journal of Human Genetics, 14, 1189–1194.Lyst, M. J., Ekiert, R., Ebert, D. H., Merusi, C., Nowak, J., Selfridge, J., et al. (2013). Rett syndrome mutations abolish the interaction of MeCP2 with the NCoR/SMRT

co-repressor. Nature Neuroscience, 7, 898–902.Mellen, M., Ayata, P., Dewell, S., Kriaucionis, S., & Heintz, N. (2012). MeCP2 binds to 5hmc enriched within active genes and accessible chromatin in the nervous

system. Cell, 151, 1417–1430.Mnatzakanian, G. N., Lohi, H., Munteanu, I., Alfred, S. E., Yamada, T., MacLeod, P. J., et al. (2004). A previously unidentified MECP2 open reading frame defines a new

protein isoform relevant to Rett syndrome. Nature Genetics, 36, 339–341.Monros, E., Armstrong, J., Aibar, E., Poo, P., Canos, I., & Pineda, M. (2001). Rett syndrome in Spain: Mutation analysis and clinical correlations. Brain Development, 23,

S251–S253.Myles, B. S., & Simpson, R. L. (1998). Asperger syndrome: A guide for educators and parents. Austin, TX: PRO-ED.Nan, X., Campoy, F. J., & Bird, A. (1997). MeCP2 is a transcriptional repressor with abundant binding sites in genomic chromatin. Cell, 88, 471–481.Nan, X., Ng, H. H., Johnson, C. A., Laherty, C. D., Turner, B. M., Eisenman, R. N., et al. (1998). Transcriptional repression by the methyl-CpG-binding protein MeCP2

involves a histone deacetylase complex. Nature, 393, 386–399.Neul, J. L., Fang, P., Barrish, J., Lane, J., Caeg, E. B., Smith, E. O., et al. (2008). Specific mutations in Methyl-CpG-Binding Protein 2 confer different severity in Rett

syndrome. Neurology, 70, 1313–1321.Nikitina, T., Shi, X., Ghosh, R. P., Horowitz-Scherer, R. A., Hansen, J. C., & Woodcock, C. L. (2007). Multiple modes of interaction between the methylated DNA

binding protein MeCP2 and chromatin. Molecular and Cellular Biology, 27, 864–877.Percy, A. K., Lane, J. B., Childers, J., Skinner, S., Annese, F., Barrish, J., et al. (2007). Rett syndrome: North American database. Journal of Child Neurology, 22,

1338–1341.Renieri, A., Mari, F., Mencarelli, M. A., Scala, E., Ariani, F., Longo, I., et al. (2009). Diagnostic criteria for the Zappella variant of Rett syndrome (the preserved speech

variant). Brain and Development, 31, 208–216.Scala, E., Ariani, F., Mari, F., Caselli, R., Pescussi, C., Longo, I., et al. (2005). CDKL5/STK9 is mutated in Rett syndrome variant with infantile spasms. Journal of Medical

Genetics, 42, 103–107.Schopler, E., Reichler, R., & Rochen-Renner, B. (1998). Childhood autism rating scale. Los Angeles, CA: Western Psychological Services.Smeets, E. E. J., Pelc, K., & Dan, B. (2011). Rett syndrome. Molecular Syndromology, 2, 113–127.

Page 11: Recent insights into genotype-phenotype relationships in patients with Rett syndrome using a fine grain scale

R.A. Fabio et al. / Research in Developmental Disabilities 35 (2014) 2976–29862986

Van den Veyver, I. B., & Zoghbi, H. Y. (2001). Mutations in the gene encoding methyl-CpG-binding protein 2 cause Rett syndrome. Brain Development, 23,S147–S151.

Vignoli, A., Fabio, R. A., La Briola, F., Giannatiempo, S., Antonietti, A., Maggiolini, S., et al. (2010). Correlations between neurophysiological behavioral and cognitivefunction in Rett syndrome. Epilepsy and Behaviour, 17, 489–496.

Weaving, L. S., Williamson, S. L., Bennets, B., Davis, M., Ellaway, C. J., Leonard, H., et al. (2003). Effects of MECP2 mutation type, location and X-inactivation inmodulating Rett syndrome phenotype. American Journal of Medical Genetics – Part A, 118, 103–114.

Weng, S. M., Bailey, M. E. S., & Cobb, S. R. (2011). Rett syndrome: From bed to bench. Pediatrics and Neonatology, 52, 309–316.