Skip to main content
Erschienen in: BMC Dermatology 1/2020

Open Access 01.12.2020 | Research article

Mapping of cis-acting expression quantitative trait loci in human scalp hair follicles

verfasst von: Marisol Herrera-Rivero, Lara M. Hochfeld, Sugirthan Sivalingam, Markus M. Nöthen, Stefanie Heilmann-Heimbach

Erschienen in: BMC Dermatology | Ausgabe 1/2020

Abstract

Background

The association of molecular phenotypes, such as gene transcript levels, with human common genetic variation can help to improve our understanding of interindividual variability of tissue-specific gene regulation and its implications for disease.

Methods

With the aim to capture the spectrum of biological processes affected by regulatory common genetic variants (minor allele frequency ≥ 1%) in healthy hair follicles (HFs) from scalp tissue, we performed a genome-wide mapping of cis-acting expression quantitative trait loci (eQTLs) in plucked HFs, and applied these eQTLs to help further explain genomic findings for hair-related traits.

Results

We report 374 high-confidence eQTLs found in occipital scalp tissue, whose associated genes (eGenes) showed enrichments for metabolic, mitotic and immune processes, as well as responses to steroid hormones. We were able to replicate 68 of these associations in a smaller, independent dataset, in either frontal and/or occipital scalp tissue. Furthermore, we found three genomic regions overlapping reported genetic loci for hair shape and hair color. We found evidence to confirm the contributions of PADI3 to human variation in hair traits and suggest a novel potential candidate gene within known loci for androgenetic alopecia.

Conclusions

Our study shows that an array of basic cellular functions relevant for hair growth are genetically regulated within the HF, and can be applied to aid the interpretation of interindividual variability on hair traits, as well as genetic findings for common hair disorders.
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1186/​s12895-020-00113-y.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
AGA
Androgenetic alopecia
eQTLs
Expression quantitative trait loci
FDR
False discovery rate
FUMA
Functional mapping and annotation of GWAS
GWAS
Genome-wide association analysis
HF(s)
Hair follicle(s)
HDACs
Histone deacetylases
IFN
Interferon
LD
Linkage disequilibrium
MAF
Minor allele frequency
PCA
Principal component analysis
PCs
Principal components
RIN
RNA integrity number
SNP
Single nucleotide polymorphism
TSS
Transcription start site

Background

Human hair traits show wide interindividual variability, which has been suggested to be largely determined by genetic factors [1, 2].
Systematic gene identification efforts for a growing number of quantitative and complex human traits have shown that the majority of associated genetic factors are located in non-coding genomic regions [3]. These variants most probably exert their functional effects through the tissue-specific modulation of the expression of trait-relevant genes. Expression quantitative trait loci (eQTL) analyses, that correlate sequence variation with gene expression data, have proven a valuable tool in terms of delineating the tissue-specific architecture(s) of gene regulation and predicting the impact that trait-associated variants exert on it [4]. This approach therefore bridges the gap between a genetic association finding and the underlying biological mechanism, and may provide crucial insights into disease development.
Increased knowledge of the genetic factors that contribute to variability of gene regulation in the human scalp hair follicle (HF) will aid the interpretation of genetic findings for hair-related traits and hair loss disorders, such as androgenetic alopecia (AGA). Moreover, the comparison of the regulatory architecture between HFs from different scalp areas may aid the understanding of the variable susceptibility of hair follicle subpopulations to hormonal hair loss. The aims of the present study were to: 1) perform a systematic mapping of eQTLs in the human hair follicle, and 2) evaluate the potential of these eQTLs in terms of the functional annotation of genetic loci that contribute to the development of hair-related traits and common diseases.

Methods

Sample collection

About 50 HFs were plucked from the occipital scalp of 100 (discovery sample), and the frontal and occipital scalp of 25 (replication sample, previously described in [5]) unrelated male volunteers. Peripheral venous blood samples were collected from all study participants. All volunteers were German residents of European descent, and showed a collective mean age of 27.9 years.

Extraction of nucleic acids

DNA was extracted from whole blood samples using the Chemagic Magnetic Separation Module I (Perkin Elmer Chemagen Technology Inc., Baesweiler, Germany). Total RNA was extracted from HFs using the RNeasy Micro Kit (Qiagen, Hilden, Germany), and the quality and quantity of the RNA were assessed using a BioAnalyzer 2100 (Agilent Technologies, Waldbronn, Germany), and a NanoDrop ND-1000 spectrophotometer (Peqlab Biotechnologie, Erlangen, Germany), respectively. Only total RNA samples with an RNA integrity number (RIN) ≥ 8 were further analyzed in the study.

Array hybridization

DNA extracts were hybridized onto the Human OmniExpress-12v1.0 bead array (Illumina, San Diego, CA, USA) (N = 100) or the Illumina PsychArray v1.0 (N = 25) for genome-wide genotyping, while RNA extracts from HFs were amplified and biotinylated using the TotalPrep™-96 RNA Amplification Kit (Illumina, San Diego, CA, USA) prior to whole transcriptome profiling, performed on the Illumina HT-12v4 bead array.

Preparation of genotype data

SNP array raw data was initially analyzed using the Genotyping module within the GenomeStudio software (Illumina). Genotype calls were exported for basic quality control in PLINK v1.9 ([6]; www.​cog-genomics.​org/​plink/​1.​9/​) to eliminate bad quality data (e.g. SNPs and individuals with high degree of missing data, very rare SNPs) prior to genotype imputation. Imputation was performed on the Michigan Imputation Server [7] using the 1000 Genomes Project Phase 3 v5 reference panel and Eagle v2.3 phasing [8]. Post-imputation data processing was performed using VCFtools [9] and quality control was carried out in PLINK 1.9. Briefly, only biallelic single nucleotide variants with high imputation quality score (Rsq) > 0.7, minor allele frequency (MAF) ≥ 1% and under Hardy-Weinberg equilibrium (HWE p > 1 × 10− 8) were further analyzed. A principal component analysis (PCA) was performed with PLINK to identify potential outlier samples and use the generated principal components (PCs) as covariates for eQTL analysis. Two final autosomal genotype datasets consisted of 5,887,234 SNPs and 98 individuals for the discovery sample, and 1,044,566 SNPs and 24 individuals for the replication sample.

Preparation of gene expression data

Raw data from the expression microarrays was initially analyzed using the Gene Expression module within the GenomeStudio software to generate calls and detection p-values. The probe-level gene expression data was exported for pre-processing by background correction, quantile normalization, log2 transformation, probe quality filtering and identification of potential outlier samples by PCA using R. Probes were considered expressed when showing a detection p-value < 0.01 in at least 5% of the samples. Probe quality filtering included the retention of only “good” and “perfect” quality probes mapping to only one gene with a valid identifier, according to annotations retrieved from the illuminaHumanv4.db package [10]. Three final gene expression datasets consisted of 13,217 probes and 98 individuals for the discovery sample, and 13,091 probes from frontal scalp and 12,814 probes from occipital scalp and 24 individuals for the replication sample.

eQTL mapping

Genome-wide associations in cis (1 Mb window) between the expression levels in scalp HFs and SNP genotypes were tested using QTLtools [11]. Three covariate files prepared for the analyses included the first 10 (discovery sample) or 5 (replication sample) PCs for the genotype datasets and the first 10 or 5 PCs for the phenotype datasets. Initially, the full spectrum of eQTLs for each of all three gene expression datasets was identified through nominal pass analysis. After exploration of the nominal significant results (p < 0.05), only those eQTLs with false discovery rate (FDR) < 1 × 10− 4 in the discovery sample were considered true eQTLs, while all eQTLs with p < 0.01 in the replication sample were retained for further analyses. To identify independent signals within our set of true eQTLs, a permutation pass analysis (1000 permutations), followed by a conditional pass analysis on grouped phenotypes (i.e. a gene-level output from the probe-level analysis) were applied.

Variant annotation

We explored reported eQTL effects for our independent HF eQTLs using the Variant Annotation tool from SNiPA (Single Nucleotide Polymorphisms Annotator) ([12]; http://​snipa.​org), noting whether or not the HF eQTL has been previously reported to have cis-eQTL effects in at least one tissue, and whether or not eQTL effects have been observed on the same gene as in our study. Additionally, we searched for trait associations of our true HF eQTLs that have been reported in the GWAS Catalog ([13]; https://​www.​ebi.​ac.​uk/​gwas/​).

Replication of HF eQTLs

The true eQTLs identified from the nominal pass analysis in the discovery sample were considered replicated when the same eQTL SNP (eSNP) was found associated to the same eGene, either through the same or a different probe, with p < 0.01 in either frontal and/or occipital scalp areas from the replication sample.

Differential eQTLs between frontal and occipital scalp

To investigate differential regulatory effects between frontal and occipital scalp areas, non-overlapping eQTLs with p-value < 5 × 10− 5 were investigated in the replication sample. These included eSNPs that showed different effects between both datasets (i.e. association to a different eGene or opposing direction of the effect for the same SNP-gene pair; different-effect eQTLs), and those eQTLs that were unique to the frontal or occipital dataset (i.e. no overlaps in eSNPs or eGenes; region-specific eQTLs).

Functional enrichment analysis

The lists of eGenes obtained from the discovery sample and the differential eQTL analysis were submitted for enrichment analysis using the GENE2FUNC function of the Functional Mapping and Annotation of Genome-Wide Association Studies (FUMA GWAS) platform ([14]; http://​fuma.​ctglab.​nl/​). Each analysis used the provided option to include all human genes as background, a significance cutoff of FDR < 0.05 and a threshold of minimum overlapping genes with the gene-sets = 3. In addition, a comparative analysis of biological pathways for the differential eQTLs was performed using the lists of the identified region-specific eGenes and the Gene Enrichment Compare function within the FunRich (Functional Enrichment analysis tool) software [15]. From the resulting pathway analysis using the default FunRich database, only frontal- and occipital-specific terms with at least 2 genes from the dataset present in the pathway term, and p < 0.01 (from hypergeometric test) in one scalp area and p > 0.05 in the other, were retained for the purposes of the present study.

Overlaps with reported genetic findings for hair phenotypes

To test the informativeness of our HF eQTLs for the interpretation of genetic findings for hair phenotypes, we used three sets of published genome-wide significant variants (p < 5 × 10− 8) associated with (i) hair shape (discovery meta-analysis, supplementary table 2 from [1]), (ii) hair color (meta-analysis, supplementary table 2 from [2]), and (iii) AGA (reported lead SNPs from [1626]). These three variant sets were subjected to analysis by the SNP2GENE function of FUMA GWAS to assign GWAS findings to genomic regions. The analysis for each phenotype was set to include SNPs from the 1000 Genomes Project (Phase 3, European population) [27] that are in linkage disequilibrium (LD) with the reported GWAS variants. LD blocks were set to include variants of minor allele frequency ≥ 0.01, r2 ≥ 0.6 and a distance < 500 kb for merging into a locus. Afterwards, we searched for overlaps between our true HF eQTLs and the resulting loci for hair shape, hair color and AGA.

Results

Hair follicle eQTLs

We identified almost 3 million nominally significant eQTL signals in HFs from the occipital scalp (Fig. 1a). However, when we plotted the p-values and distances to the transcription start site (TSS) of the probe for each top SNP-probe pair with FDR < 0.05, we observed that a large number of SNPs were lying over 250 kb from the probe’s TSS, suggesting many of these might actually be noise signals (false positives). For this reason, only eQTL-findings that showed an FDR < 1 × 10− 4 were considered the likely true eQTLs (14,497 probe-SNP pairs) (Fig. 1b). From this set of true eQTLs, we identified a total of 374 independent HF eQTLs (Fig. 2a, Suppl.Table.1 [Supplementary Tables]). The top 20 independent eQTL findings are presented in Table 1. Functional enrichment analyses using FUMA revealed that associated eGenes were highly enriched (FDR < 0.05) for a variety of Hallmark and Reactome pathways related to metabolism, immune functions, cellular proliferation, apoptosis, adipogenesis and responses to sex hormones (Fig. 1c, Suppl.Table.2 [Supplementary Tables]). Additionally, the annotation of reported eQTL effects with SNiPA showed that 266 of 374 independent eSNPs have been previously reported to have cis-eQTL effects in at least one other tissue, from which 228 were reported to affect the same eGene. Annotation of GWAS Catalog trait associations identified 13 independent HF eSNPs that are associated with 15 traits, including type 2 diabetes mellitus and chronic inflammatory diseases. From the set of true HF eQTLs, 185 eSNPs showed associations with 134 traits, with the highest number of associations found for body mass index (associated with 29 eSNPs) and blood protein levels (associated with 13 eSNPs) (data not shown).
Table 1
Top findings for hair follicle eQTLs
Gene
# eQTLs
Best eSNP
Chr
Position
A1
A2
P
FDR
Effect
Top 20 independent eQTLs (discovery sample)
IPO8
183
rs7326
12
30,782,301
G
A
5.17E-31
1.56E-24
−0.6840
ATP5MD
95
rs2271751
10
105,175,131
C
T
5.80E-26
3.21E-20
1.1004
C17ORF97
35
rs11150881
17
259,304
G
A
2.18E-25
1.18E-19
0.6600
TIMM10
9
rs3851117
11
57,237,113
A
G
4.60E-25
2.11E-19
−0.8391
KCTD10
211
rs4766601
12
109,890,080
C
G
9.89E-25
4.49E-19
0.5773
C5ORF35/SETD9
50
rs2591963
5
56,237,135
A
T
1.50E-24
6.43E-19
−0.6930
NDUFS5
45
rs374279960
1
39,484,742
G
T
2.36E-24
1.01E-18
−0.2496
LOC339804
267
rs3213944
2
61,372,298
G
C
8.36E-24
2.50E-18
0.4471
FMO1
45
rs6674596
1
171,235,088
T
A
1.04E-23
3.07E-18
1.3913
PARP12
32
rs2286197
7
139,726,467
G
A
1.19E-23
3.46E-18
0.3402
SH3YL1
50
rs62114506
2
242,793
C
G
1.30E-23
3.78E-18
−0.5886
PEX6
57
rs3805946
6
42,955,749
C
T
1.47E-23
4.11E-18
0.5651
FN3KRP
109
rs2249888
17
80,675,738
G
A
1.54E-23
4.11E-18
0.3341
BCR
44
rs131703
22
23,652,201
G
A
6.23E-23
1.36E-17
0.4445
PSMG1
295
rs35064900
21
40,555,492
G
A
7.92E-23
1.55E-17
−0.3917
RPS26L
47
rs773114
12
56,379,060
T
A
1.14E-22
2.12E-17
0.4386
GOLGB1
24
rs9968051
3
121,384,081
G
C
3.20E-22
5.56E-17
−0.6104
ERP27
182
rs12312821
12
15,077,527
T
A
4.48E-22
7.63E-17
−0.6385
SLC47A2
51
rs12451902
17
19,619,063
G
A
1.02E-21
1.65E-16
0.6855
ABHD12
517
rs2258728
20
25,276,343
A
G
1.91E-21
2.88E-16
0.2737
Top 10 eQTLs in frontal scalp (replication sample)
TDRD6
14
rs552053483
6
46,870,456
C
A
5.32E-12
4.08E-06
0.4313
CXCL1
32
rs191105830
4
75,248,687
G
A
3.07E-09
7.16E-04
0.4703
ZXDC
12
rs552704609
3
126,534,497
A
C
4.70E-08
0.00871
0.3945
COA4
2
rs34214542
11
74,207,014
G
T
7.82E-08
0.01141
0.3317
DLG2
25
rs141374880
11
83,245,973
A
G
8.60E-08
0.01141
0.4126
TADA2A
3
rs74398459
17
36,095,246
G
T
1.17E-07
0.01446
0.3338
C17orf67
2
rs187188078
17
55,455,184
T
C
1.46E-07
0.01761
0.6551
FAM220A
16
rs188128204
7
6,373,651
A
G
2.11E-07
0.02083
0.4342
ARMC6
5
rs150042710
19
18,518,809
T
C
2.40E-07
0.02262
0.6410
NUP153
10
rs115808997
6
17,571,313
T
C
4.54E-07
0.03605
−0.4850
Top 10 eQTLs in occipital scalp (replication sample)
TELO2
17
rs561125041
16
1,464,247
T
C
8.51E-10
3.00E-04
1.2115
LPHN1
17
rs147147550
19
14,373,177
T
C
9.90E-10
3.00E-04
0.6231
ENTHD2
5
rs117860115
17
78,992,225
T
C
4.10E-09
0.00121
0.1629
GCSAM
2
rs139726887
3
112,327,159
T
C
3.41E-08
0.00496
0.3622
CCL2
11
rs72825069
17
32,600,928
T
C
6.38E-08
0.00678
1.7221
RUFY3
3
rs115702490
4
72,452,170
T
C
1.24E-07
0.01296
0.4149
GGPS1
6
rs115335216
1
236,067,212
T
G
2.13E-07
0.01422
0.2215
MTF1
10
rs3935450
1
37,656,324
G
T
2.65E-07
0.01763
−0.2218
CXADR
4
rs55707799
21
18,849,240
C
T
3.47E-07
0.02132
−0.4272
KRT37
1
rs35490951
17
40,250,939
C
A
3.65E-07
0.02132
1.8106
The top 20 hair follicle eQTLs obtained from human occipital scalp in the discovery sample and the top 10 hair follicle eQTLs obtained from frontal and occipital scalp in the replication sample are shown
eQTLs expression quantitative trait loci, eSNP SNP with eQTL effect, Chr chromosome, A1 effect allele, A2 other allele, P p-value, FDR false discovery rate
We achieved the replication of 68 independent signals (Fig. 2) in either occipital and/or frontal scalp HFs from a total of 255 overlapping eQTLs (Suppl.Table.3 [Supplementary Tables]). In total, 188 and 157 eSNPs, and 264 and 276 eGenes, overlapped between our true eQTLs and the frontal and occipital HF eQTLs from the replication sample, respectively. Although only 92 eSNPs overlapped between all three eQTL datasets, these accounted for 249 eGenes (Fig. 1d and e); however, not all eQTL effects were consistent between the three datasets.

Differential HF eQTLs between frontal and occipital scalp areas

We included all eQTLs with p < 5 × 10− 5 in the analysis of the replication sample. However, as the power to detect associations is markedly reduced due to the limited sample size (N = 24), this is reflected in the number and statistical significance of eQTLs that we detected in frontal and occipital scalp, as well as in their distributions with respect to the probe’s TSS (Suppl.Figure.1 [Supplementary Figures]). In general, we found little overlap between these two datasets, and will limit ourselves here to briefly present an overview of regional differences between frontal and occipital scalp.
We identified 71 HF eQTLs with inconsistent effects between frontal and occipital HFs. These were considered “different-effect eQTLs” and affected 11 genes (Suppl.Table.4 [Supplementary Tables]). Furthermore, we identified 289 frontal (Suppl.Table.5 [Supplementary Tables]) and 339 (Suppl.Table.6 [Supplementary Tables]) occipital (region-specific) eQTLs, from which the top 10 for each scalp region are shown in Table 1. To identify differentially-enriched pathways, we performed a separate FUMA pathway analysis for frontal and occipital differential eGenes (i.e. region-specific + different-effect) and compared the results for both scalp regions. This analysis suggested differences in the genetically-determined regulation of the responses to steroid hormones, cell cycle control, cellular metabolism and immune functions between the scalp regions (Suppl. Figure.2 [Supplementary figures]). However, as several of the region-specific pathway terms seemed redundant, we sought to further elucidate the differences between both scalp regions through a comparative pathway analysis using FunRich. This analysis suggested that the more important differences between frontal and occipital scalp HFs are on pathways related to the metabolism of amino acids and signaling by histone deacetylases (HDACs) in the frontal scalp, and to proliferative processes in the occipital area (Table 2).
Table 2
Best frontal- and occipital-specific biological pathways enriched for differential eGenes
Biological pathway
GeneSet
Frontal scalp
Occipital scalp
n
Fold value
P
n
Fold value
P
Frontal scalp eQTLs
 Activation of Chaperones by ATF6-alpha
9
3
21.85
2.78E-04
0
0.01
1
 Interconversion of 2-oxoglutarate and 2-hydroxyglutarate
3
2
43.68
6.92E-04
0
0.01
1
 Pyruvate metabolism and Citric Acid (TCA) cycle
31
4
8.46
1.22E-03
0
0.01
1
 Arginine degradation I (arginase pathway)
4
2
32.78
1.37E-03
0
0.01
1
 Proline biosynthesis II (from arginine)
5
2
26.24
2.26E-03
0
0.01
1
 Arginine degradation VI (arginase 2 pathway)
6
2
21.87
3.36E-03
0
0.01
1
 Vitamin C (ascorbate) metabolism
6
2
21.87
3.36E-03
0
0.01
1
 Dolichyl-diphosphooligosaccharide biosynthesis
7
2
18.75
4.65E-03
0
0.01
1
 Signaling events mediated by HDAC Class I
113
6
3.48
7.84E-03
1
0.51
8.66E-01
 Signaling events mediated by HDAC Class II
38
4
6.90
2.62E-03
1
1.51
4.91E-01
 BMP receptor signaling
226
9
2.61
8.19E-03
5
1.26
3.66E-01
 Amino acid synthesis and interconversion (transamination)
10
3
19.67
3.92E-04
1
5.74
1.63E-01
Occipital scalp eQTLs
 BARD1 signaling events
29
1
2.28
3.61E-01
4
7.86
1.57E-03
 G0 and Early G1
21
0
0.01
1
3
8.15
5.66E-03
 TGFBR
125
4
2.10
1.25E-01
7
3.19
6.64E-03
 VEGF and VEGFR signaling network
1301
27
1.36
6.57E-02
35
1.53
8.04E-03
eQTLs expression quantitative trait loci, GeneSet number of total genes in pathway term, n number of genes from the study overlapping the pathway term, P p-vlaue

Overlaps with genomic regions associated with hair phenotypes

To illustrate the applicability of our set of true eQTLs to interpret GWAS findings for hair phenotypes, we investigated for a potential overlap of HF eSNPs with genomic risk loci for AGA (98 loci), hair shape (12 loci) and hair color (122 loci) (Table 3). We identified one genomic locus at 1p36.13 that had previously been associated with hair phenotypes. This comprised 31 eSNPs overlapping association signals for hair shape (overlapping lead SNP: rs11203346) and hair color (overlapping lead SNP: rs72646785) (Fig. 3a). For this locus, the nearest genes were mapped to PADI3 and PADI4 in both GWAS, while the eGene for this region corresponded to PADI3. Moreover, another locus at 4q21.21, comprised by 26 eSNPs linked to the signal of rs6533756 (a non-eSNP in LD r2 ≥ 0.8 with all eSNPs), was associated with hair shape. The overlapping eSNP most strongly associated with hair shape at this locus was rs7695038 (p = 3.9 × 10− 10). While eSNPs at this locus locate within a region coding for the FRAS1 gene, the associated eGene, ANXA3, was located 66,182 bp downstream of FRAS1 (Fig. 3b). A third locus at 17q25.3 comprised the eSNP, rs8070929, which was in LD (r2 ≥ 0.97) with rs34872037 (a non-eSNP) that is associated with hair color. While the nearest gene for the SNP at this locus was NPLOC4, the HF eGene corresponded to TSPAN10 (Fig. 3c). All eSNPs overlapping association signals for hair shape and/or hair color can be found in Suppl.Table.7. We found no overlaps of our true HF eQTL loci with the reported genome-wide significant findings for AGA. However, we observed that the eGene ATP2B4 was annotated as the nearest gene to an AGA genomic locus at 1q32.1. Moreover, it is perhaps worth to notice that, at the nominal significance level (p < 0.05), we found 93 eGenes to overlap with reported AGA candidate genes, of which 18 were associated with eSNPs at an FDR < 0.05 (ALPL, B3GNT8, BCL2, DIP2B, FAM136A, HDAC9, LCLAT1, LHPP, NSF, PRRX1, RPTN, RSPO2, SUCNR1, TCHH, TMEM50A, TWIST2, MSIG2 and SPAG17).
Table 3
Genomic loci for hair phenotypes and their observed overlaps with our hair follicle eQTLs
Summary
Hair shape
Hair color
AGA
Sample size
16,763
290,891
145,000a
# Genomic loci
12
122
98
# Lead SNPs
23
124
185
# Ind.Sig.SNPs
63
130
245
# Candidate SNPs
1424
6658
11,821
# GWAS SNPs
704
134
335
# Mapped genes
69
204
161
# Total input SNPs
706
137
387
# Skipped input SNPs
2
3
36
Overlapping eGenes with mapped genes
PADI3
BCAS1, C17orf70, CHMP4C, MRPL39, PADI3, RYBP
ATP2B4
Overlapping loci
2
2
Overlapping Chr.Loc.
chr1p36.13
chr1p36.13
chr4q21.21
chr17q25.3
 
Mapped genes
PADI3, PADI4
PADI3, PADI4
FRAS1
NPLOC4
 
eGenes
PADI3
PADI3
ANXA3
TSPAN10
 
AGA androgenetic alopecia, Ind.Sig.SNPs independent significant SNPs, Chr.Loc. chromosome location
aApproximated sample size, as lead SNPs were collected from different studies

Discussion

The strongest eQTL associations in occipital HFs were observed for IPO8 (rs7326), ATP5MD (rs2271751), and C17orf97/LIAT1 (rs11150881). The findings for ATP5MD were confirmed in our small replication study. Although we were not able to replicate the “true” SNP-gene associations for IPO8 in our replication sample, a set of different SNPs were indeed associated with IPO8 expression in HFs from frontal and/or occipital scalp areas, some of which were also associated with IPO8 expression in the discovery sample at the nominal level (data not shown), therefore confirming the genetic regulation of IPO8 in HFs. While little is known about the function of C17orf97, IPO8 mediates the nuclear import of proteins and mature microRNAs [28]. ATP5MD is crucial for the maintenance of ATP synthase in mitochondria, and might actively participate in the cellular energy metabolism, a process with well-known relevance to hair biology and hair growth [29, 30]. Over-expression of ATP5MD causes a number of mitochondrial abnormalities and an increase in anaerobic metabolism associated with the induction of an epithelial to mesenchymal-like transition, as well as delayed cell growth [31, 32]. The genetically-controlled regulation of mitochondrial function in human HF is further supported by our pathway analysis that found an enrichment of eGenes for true eQTLs in pathways related to mitochondrial function along with other pathways, such as the regulation of responses to steroid hormones, the Wnt/β-catenin and interferon (IFN) signaling, adipogenesis, immune responses and the metabolism of glucose and lipids, all of which have well-known roles in HF biology [33].
We also investigated whether there might be differences in the genetic control of gene expression between HF subpopulations from different scalp areas. Despite the reduced size of this (replication) sample, our results point to interesting avenues for future research. For instance, an important difference between frontal and occipital scalp appears to be the metabolism of amino acids, including arginine degradation. It is known that the HF is dependent on arginine, as hair growth depends on the vasculature and L-arginine not only participates in cell proliferation but is a precursor for the vascular mediator nitric oxide. L-arginine deficiency has been shown to impair hair elongation, while its supplementation increases the number of HFs in anagen (growth) and decreases that of HFs in telogen (resting) phase [33, 34]. This supports the notion that regional increases in arginine degradation might result in enhanced vulnerability to hair loss in the frontal scalp area. A similar scenario can be thought for HDACs. HDACs are important transcriptional repressors that act in multiprotein complexes and are involved in the control of cell cycle progression [35, 36]. While class I HDACs are ubiquitously expressed, class II HDACs show tissue specificity [36]. In our study, we found eQTLs for HDACs 2, 5 and 7 only in frontal scalp. Interestingly, these particular HDACs modulate HF development and homeostasis [37, 38], as well as angiogenesis and vascular integrity [3941]. Moreover, androgen actions have been shown to regulate HDAC7 subcellular compartmentalization, and HDAC7 has been proposed to be a co-repressor of the androgen receptor [42].
We also found evidence for differential regulation of vitamin C (L-ascorbic acid) metabolism in frontal scalp. It has been shown that a derivative of L-ascorbic acid (L-ascorbic acid 2-phosphate) promotes HF growth that is mediated by the induced expression of insulin-like growth factor-1 (IGF-1) in dermal papilla cells [43]. This opens the possibility that decreased expression of genes involved in the metabolism of L-ascorbic acid in frontal scalp might render this region more susceptible to hair loss. Taken together, our pathway analysis results show that frontal-specific pathways present several factors with negative effects on hair growth (e.g. androgen and estrogen responses, BMP2/4 signaling, IFN-γ response, endocannabinoid signaling), while occipital-specific pathways are more consistent with factors exerting positive effects on hair growth (e.g. vascular endothelial growth factor signaling, transforming growth factor beta receptor), considering what Bernard has referred to as “the Yin Yang of the human hair follicle” [33]. Nevertheless, the implications of genetic regulation in frontal scalp for hair loss disorders should be investigated through the generation of a confident eQTL dataset with increased power in future studies.
With our study, we also show that tissue-specific eQTL data are a valuable resource to identify regulatory effects at disease- and trait-associated loci. In particular, our results suggest an important role of PADI3 in hair traits. Indeed, PADI3 is located in the inner root sheath and medulla in anagen HFs and has been reported to play roles in HF differentiation [44] and hair shaft formation [45]. Our results also implicate ANXA3 and TSPAN10 as novel candidate genes for hair shape and color, respectively. Although we found no overlap between the true HF eQTLs and AGA genetic risk loci, which might be an expected finding, considering that the occipital scalp is not susceptible to balding, the overlapping eGene ATP2B4 provides a potential novel candidate gene for AGA, as the reported gene for the region is so far SOX13 [25, 26].
The most obvious limitation of our study is the sample size. However, we tried to overcome this limitation by applying high quality standards to the data and stringent selection criteria to our HF eQTL results. Another potential limitation of our study resides in the use of plucked HFs instead of intact HFs. However, a skin biopsy is required in order to obtain intact HFs, whereas hair plucking is a less invasive technique. Moreover, it has been demonstrated that plucked hairs retain most epithelial structures, maintain the integrity of the outer root sheet and also contain stem cells [46]. Finally, due to the small size of our samples, particularly that of the replication sample, we considered advisable to exclude from the present study indels and sex chromosomes, and limit ourselves to reporting, in very general terms, differential findings between HFs from frontal and occipital scalp areas.

Conclusions

Our analyses demonstrate that well-established HF molecular pathways are genetically regulated and that, to some extent, this regulation can show regional specificity within scalp HFs. The enriched pathways mainly underscored processes that relate to hair growth and the HF cycle. More detailed tissue-specific analyses will be enabled by future increases in sample size, ensuring an improved understanding of the genetically determined variability in HF gene expression and its implications for hair-related traits and hair loss disorders.

Acknowledgments

The authors thank the study participants for their cooperation. MMN and MHR are members of the Excellence Cluster ImmunoSensation.
All participants provided written informed consent. The study was approved by the Institutional Ethics Review Committee of the University of Bonn and experiments were conducted in accordance with the Declaration of Helsinki.
‘Not applicable’

Competing interests

The authors declare that they have no competing interests.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat Liu F, Chen Y, Zhu G, Hysi PG, Wu S, Adhikari K, et al. Meta-analysis of genome-wide association studies identifies 8 novel loci involved in shape variation of human head hair. Hum Mol Genet. 2018;27(3):559–75.PubMedCrossRef Liu F, Chen Y, Zhu G, Hysi PG, Wu S, Adhikari K, et al. Meta-analysis of genome-wide association studies identifies 8 novel loci involved in shape variation of human head hair. Hum Mol Genet. 2018;27(3):559–75.PubMedCrossRef
2.
Zurück zum Zitat Hysi PG, Valdes AM, Liu F, Furlotte NA, Evans DM, Bataille V, et al. Genome-wide association meta-analysis of individuals of European ancestry identifies new loci explaining a substantial fraction of hair color variation and heritability. Nat Genet. 2018;50(5):652–6.PubMedPubMedCentralCrossRef Hysi PG, Valdes AM, Liu F, Furlotte NA, Evans DM, Bataille V, et al. Genome-wide association meta-analysis of individuals of European ancestry identifies new loci explaining a substantial fraction of hair color variation and heritability. Nat Genet. 2018;50(5):652–6.PubMedPubMedCentralCrossRef
3.
4.
5.
Zurück zum Zitat Hochfeld LM, Anhalt T, Reinbold CS, Herrera-Rivero M, Fricker N, Nöthen MM, Heilmann-Heimbach S. Expression profiling and bioinformatic analyses suggest new target genes and pathways for human hair follicle related microRNAs. BMC Dermatol. 2017;17(1):3.PubMedPubMedCentralCrossRef Hochfeld LM, Anhalt T, Reinbold CS, Herrera-Rivero M, Fricker N, Nöthen MM, Heilmann-Heimbach S. Expression profiling and bioinformatic analyses suggest new target genes and pathways for human hair follicle related microRNAs. BMC Dermatol. 2017;17(1):3.PubMedPubMedCentralCrossRef
6.
Zurück zum Zitat Chang CC, Chow CC, Tellier LC, Vattikuti S, Purcell SM, Lee JJ. Second-generation PLINK: rising to the challenge of larger and richer datasets. Gigascience. 2015;4:7.PubMedPubMedCentralCrossRef Chang CC, Chow CC, Tellier LC, Vattikuti S, Purcell SM, Lee JJ. Second-generation PLINK: rising to the challenge of larger and richer datasets. Gigascience. 2015;4:7.PubMedPubMedCentralCrossRef
7.
Zurück zum Zitat Das S, Forer L, Schönherr S, Sidore C, Locke AE, Kwong A, et al. Next-generation genotype imputation service and methods. Nat Genet. 2016;48(10):1284–7.PubMedPubMedCentralCrossRef Das S, Forer L, Schönherr S, Sidore C, Locke AE, Kwong A, et al. Next-generation genotype imputation service and methods. Nat Genet. 2016;48(10):1284–7.PubMedPubMedCentralCrossRef
8.
Zurück zum Zitat Loh PR, Danecek P, Palamara PF, Fuchsberger C, Reshef AY, Finucane KH, et al. Reference-based phasing using the haplotype reference consortium panel. Nat Genet. 2016;48(11):1443–8.PubMedPubMedCentralCrossRef Loh PR, Danecek P, Palamara PF, Fuchsberger C, Reshef AY, Finucane KH, et al. Reference-based phasing using the haplotype reference consortium panel. Nat Genet. 2016;48(11):1443–8.PubMedPubMedCentralCrossRef
9.
Zurück zum Zitat Danecek P, Auton A, Abecasis G, Albers CA, Banks E, DePristo MA, et al. The variant call format and VCFtools. Bioinformatics. 2011;27(15):2156–8.PubMedPubMedCentralCrossRef Danecek P, Auton A, Abecasis G, Albers CA, Banks E, DePristo MA, et al. The variant call format and VCFtools. Bioinformatics. 2011;27(15):2156–8.PubMedPubMedCentralCrossRef
10.
Zurück zum Zitat Dunning M, Lynch A, Eldridge M. illuminaHumanv4.db: Illumina HumanHT12v4 annotation data (chip illuminaHumanv4). R package version 1.26.0; 2015. Dunning M, Lynch A, Eldridge M. illuminaHumanv4.db: Illumina HumanHT12v4 annotation data (chip illuminaHumanv4). R package version 1.26.0; 2015.
11.
Zurück zum Zitat Delaneau O, Ongen H, Brown AA, Fort A, Panousis NI, Dermitzakis ET. A complete tool set for molecular QTL discovery and analysis. Nat Commun. 2017;8:15452.PubMedPubMedCentralCrossRef Delaneau O, Ongen H, Brown AA, Fort A, Panousis NI, Dermitzakis ET. A complete tool set for molecular QTL discovery and analysis. Nat Commun. 2017;8:15452.PubMedPubMedCentralCrossRef
12.
Zurück zum Zitat Arnold M, Raffler J, Pfeufer A, Suhre K, Kastenmüller G. SNiPA: an interactive, genetic variant-centered annotation browser. Bioinformatics. 2015;31(8):1334–6 Available at http://www.snipa.org. Accessed Oct 2018.PubMedCrossRef Arnold M, Raffler J, Pfeufer A, Suhre K, Kastenmüller G. SNiPA: an interactive, genetic variant-centered annotation browser. Bioinformatics. 2015;31(8):1334–6 Available at http://​www.​snipa.​org. Accessed Oct 2018.PubMedCrossRef
13.
Zurück zum Zitat MacArthur J, Bowler E, Cerezo M, Gil L, Hall P, Hastings E, et al. The new NHGRI-EBI catalog of published genome-wide association studies (GWAS catalog). Nucleic Acids Res. 2017;45(D1):D896–901.PubMedCrossRef MacArthur J, Bowler E, Cerezo M, Gil L, Hall P, Hastings E, et al. The new NHGRI-EBI catalog of published genome-wide association studies (GWAS catalog). Nucleic Acids Res. 2017;45(D1):D896–901.PubMedCrossRef
14.
Zurück zum Zitat Watanabe K, Taskesen E, van Bochoven A, Posthuma D. Functional mapping and annotation of genetic associations with FUMA. Nat Commun. 2017;8(1):1826.PubMedPubMedCentralCrossRef Watanabe K, Taskesen E, van Bochoven A, Posthuma D. Functional mapping and annotation of genetic associations with FUMA. Nat Commun. 2017;8(1):1826.PubMedPubMedCentralCrossRef
15.
Zurück zum Zitat Pathan M, Keerthikumar S, Ang CS, Gangoda L, Quek CY, Williamson NA, et al. FunRich: an open access standalone functional enrichment and interaction network analysis tool. Proteomics. 2015;15(15):2597–601.PubMedCrossRef Pathan M, Keerthikumar S, Ang CS, Gangoda L, Quek CY, Williamson NA, et al. FunRich: an open access standalone functional enrichment and interaction network analysis tool. Proteomics. 2015;15(15):2597–601.PubMedCrossRef
16.
Zurück zum Zitat Hillmer AM, Flaquer A, Hanneken S, Eigelshoven S, Kortüm AK, Brockschmidt FF, et al. Genome-wide scan and fine-mapping linkage study of androgenetic alopecia reveals a locus on chromosome 3q26. Am J Hum Genet. 2008;82(3):737–43.PubMedPubMedCentralCrossRef Hillmer AM, Flaquer A, Hanneken S, Eigelshoven S, Kortüm AK, Brockschmidt FF, et al. Genome-wide scan and fine-mapping linkage study of androgenetic alopecia reveals a locus on chromosome 3q26. Am J Hum Genet. 2008;82(3):737–43.PubMedPubMedCentralCrossRef
17.
Zurück zum Zitat Richards JB, Yuan X, Geller F, Waterworth D, Bataille V, Glass D, et al. Male-pattern baldness susceptibility locus at 20p11. Nat Genet. 2008;40(11):1282–4.PubMedPubMedCentralCrossRef Richards JB, Yuan X, Geller F, Waterworth D, Bataille V, Glass D, et al. Male-pattern baldness susceptibility locus at 20p11. Nat Genet. 2008;40(11):1282–4.PubMedPubMedCentralCrossRef
18.
Zurück zum Zitat Brockschmidt FF, Heilmann S, Ellis JA, Eigelshoven S, Hanneken S, Herold C, et al. Susceptibility variants on chromosome 7p21.1 suggest HDAC9 as a new candidate gene for male-pattern baldness. Br J Dermatol. 2011;165(6):1293–302.PubMedCrossRef Brockschmidt FF, Heilmann S, Ellis JA, Eigelshoven S, Hanneken S, Herold C, et al. Susceptibility variants on chromosome 7p21.1 suggest HDAC9 as a new candidate gene for male-pattern baldness. Br J Dermatol. 2011;165(6):1293–302.PubMedCrossRef
19.
Zurück zum Zitat Li R, Brockschmidt FF, Kiefer AK, Stefansson H, Nyholt DR, Song K, et al. Six novel susceptibility loci for early-onset androgenetic alopecia and their unexpected association with common diseases. PLoS Genet. 2012;8(5):e1002746.PubMedPubMedCentralCrossRef Li R, Brockschmidt FF, Kiefer AK, Stefansson H, Nyholt DR, Song K, et al. Six novel susceptibility loci for early-onset androgenetic alopecia and their unexpected association with common diseases. PLoS Genet. 2012;8(5):e1002746.PubMedPubMedCentralCrossRef
20.
Zurück zum Zitat Adhikari K, Fontanil T, Cal S, Mendoza-Revilla J, Fuentes-Guajardo M, Chacón-Duque JC, et al. A genome-wide association scan in admixed Latin Americans identifies loci influencing facial and scalp hair features. Nat Commun. 2016;7:10815.PubMedPubMedCentralCrossRef Adhikari K, Fontanil T, Cal S, Mendoza-Revilla J, Fuentes-Guajardo M, Chacón-Duque JC, et al. A genome-wide association scan in admixed Latin Americans identifies loci influencing facial and scalp hair features. Nat Commun. 2016;7:10815.PubMedPubMedCentralCrossRef
21.
Zurück zum Zitat Liu F, Hamer MA, Heilmann S, Herold C, Moebus S, Hofman A, et al. Prediction of male-pattern baldness from genotypes. Eur J Hum Genet. 2016;24(6):895–902.PubMedCrossRef Liu F, Hamer MA, Heilmann S, Herold C, Moebus S, Hofman A, et al. Prediction of male-pattern baldness from genotypes. Eur J Hum Genet. 2016;24(6):895–902.PubMedCrossRef
22.
Zurück zum Zitat Pickrell JK, Berisa T, Liu JZ, Ségurel L, Tung JY, Hinds DA. Detection and interpretation of shared genetic influences on 42 human traits. Nat Genet. 2016;48(7):709–17.PubMedPubMedCentralCrossRef Pickrell JK, Berisa T, Liu JZ, Ségurel L, Tung JY, Hinds DA. Detection and interpretation of shared genetic influences on 42 human traits. Nat Genet. 2016;48(7):709–17.PubMedPubMedCentralCrossRef
23.
Zurück zum Zitat Heilmann S, Kiefer AK, Fricker N, Drichel D, Hillmer AM, Herold C, et al. Androgenetic alopecia: identification of four genetic risk loci and evidence for the contribution of WNT signaling to its etiology. J Invest Dermatol. 2013;133(6):1489–96.PubMedCrossRef Heilmann S, Kiefer AK, Fricker N, Drichel D, Hillmer AM, Herold C, et al. Androgenetic alopecia: identification of four genetic risk loci and evidence for the contribution of WNT signaling to its etiology. J Invest Dermatol. 2013;133(6):1489–96.PubMedCrossRef
24.
Zurück zum Zitat Heilmann-Heimbach S, Herold C, Hochfeld LM, Hillmer AM, Nyholt DR, Hecker J, et al. Meta-analysis identifies novel risk loci and yields systematic insights into the biology of male-pattern baldness. Nat Commun. 2017;8:14694.PubMedPubMedCentralCrossRef Heilmann-Heimbach S, Herold C, Hochfeld LM, Hillmer AM, Nyholt DR, Hecker J, et al. Meta-analysis identifies novel risk loci and yields systematic insights into the biology of male-pattern baldness. Nat Commun. 2017;8:14694.PubMedPubMedCentralCrossRef
25.
Zurück zum Zitat Hagenaars SP, Hill WD, Harris SE, Ritchie SJ, Davies G, Liewald DC, et al. Genetic prediction of male pattern baldness. PLoS Genet. 2017;13(2):e1006594.PubMedPubMedCentralCrossRef Hagenaars SP, Hill WD, Harris SE, Ritchie SJ, Davies G, Liewald DC, et al. Genetic prediction of male pattern baldness. PLoS Genet. 2017;13(2):e1006594.PubMedPubMedCentralCrossRef
26.
Zurück zum Zitat Pirastu N, Joshi PK, de Vries PS, Cornelis MC, McKeigue PM, Keum N, et al. GWAS for male-pattern baldness identifies 71 susceptibility loci explaining 38% of the risk. Nat Commun. 2017;8(1):1584.PubMedPubMedCentralCrossRef Pirastu N, Joshi PK, de Vries PS, Cornelis MC, McKeigue PM, Keum N, et al. GWAS for male-pattern baldness identifies 71 susceptibility loci explaining 38% of the risk. Nat Commun. 2017;8(1):1584.PubMedPubMedCentralCrossRef
27.
Zurück zum Zitat 1000 Genomes Project Consortium, Auton A, Brooks LD, Durbin RM, Garrison EP, Kang HM, et al. A global reference for human genetic variation. Nature. 2015;526(7571):68–74.CrossRef 1000 Genomes Project Consortium, Auton A, Brooks LD, Durbin RM, Garrison EP, Kang HM, et al. A global reference for human genetic variation. Nature. 2015;526(7571):68–74.CrossRef
28.
Zurück zum Zitat Wei Y, Li L, Wang D, Zhang CY, Zen K. Importin 8 regulates the transport of mature microRNAs into the cell nucleus. J Biol Chem. 2014;289(15):10270–5.PubMedPubMedCentralCrossRef Wei Y, Li L, Wang D, Zhang CY, Zen K. Importin 8 regulates the transport of mature microRNAs into the cell nucleus. J Biol Chem. 2014;289(15):10270–5.PubMedPubMedCentralCrossRef
29.
Zurück zum Zitat Vidali S, Knuever J, Lerchner J, Giesen M, Bíró T, Klinger M, et al. Hypothalamic-pituitary-thyroid axis hormones stimulate mitochondrial function and biogenesis in human hair follicles. J Invest Dermatol. 2014;134(1):33–42.PubMedCrossRef Vidali S, Knuever J, Lerchner J, Giesen M, Bíró T, Klinger M, et al. Hypothalamic-pituitary-thyroid axis hormones stimulate mitochondrial function and biogenesis in human hair follicles. J Invest Dermatol. 2014;134(1):33–42.PubMedCrossRef
30.
Zurück zum Zitat Singh B, Schoeb TR, Bajpai P, Slominski A, Singh KK. Reversing wrinkled skin and hair loss in mice by restoring mitochondrial function. Cell Death Dis. 2018;9(7):735.PubMedPubMedCentralCrossRef Singh B, Schoeb TR, Bajpai P, Slominski A, Singh KK. Reversing wrinkled skin and hair loss in mice by restoring mitochondrial function. Cell Death Dis. 2018;9(7):735.PubMedPubMedCentralCrossRef
31.
Zurück zum Zitat Ohsakaya S, Fujikawa M, Hisabori T, Yoshida M. Knockdown of DAPIT (diabetes-associated protein in insulin-sensitive tissue) results in loss of ATP synthase in mitochondria. J Biol Chem. 2011;286(23):20292–6.PubMedPubMedCentralCrossRef Ohsakaya S, Fujikawa M, Hisabori T, Yoshida M. Knockdown of DAPIT (diabetes-associated protein in insulin-sensitive tissue) results in loss of ATP synthase in mitochondria. J Biol Chem. 2011;286(23):20292–6.PubMedPubMedCentralCrossRef
32.
Zurück zum Zitat Kontro H, Cannino G, Rustin P, Dufour E, Kainulainen H. DAPIT over-expression modulates glucose metabolism and cell behaviour in HEK293T cells. PLoS One. 2015;10(7):e0131990.PubMedPubMedCentralCrossRef Kontro H, Cannino G, Rustin P, Dufour E, Kainulainen H. DAPIT over-expression modulates glucose metabolism and cell behaviour in HEK293T cells. PLoS One. 2015;10(7):e0131990.PubMedPubMedCentralCrossRef
34.
Zurück zum Zitat Michelet JF, Bernard BA, Juchaux F, Michelin C, El Rawadi C, Loussouarn G, Pereira R. Importance of L-Arginine for human hair growth, 28th IFSCC Meeting Proceedings; 2014. p. 1123–8. Michelet JF, Bernard BA, Juchaux F, Michelin C, El Rawadi C, Loussouarn G, Pereira R. Importance of L-Arginine for human hair growth, 28th IFSCC Meeting Proceedings; 2014. p. 1123–8.
35.
Zurück zum Zitat Verdin E, Dequiedt F, Kasler HG. Class II histone deacetylases: versatile regulators. Trends Genet. 2003;19(5):286–93.PubMedCrossRef Verdin E, Dequiedt F, Kasler HG. Class II histone deacetylases: versatile regulators. Trends Genet. 2003;19(5):286–93.PubMedCrossRef
36.
Zurück zum Zitat Gallinari P, Di Marco S, Jones P, Pallaoro M, Steinkühler C. HDACs, histone deacetylation and gene transcription: from molecular biology to cancer therapeutics. Cell Res. 2007;17(3):195–211.PubMedCrossRef Gallinari P, Di Marco S, Jones P, Pallaoro M, Steinkühler C. HDACs, histone deacetylation and gene transcription: from molecular biology to cancer therapeutics. Cell Res. 2007;17(3):195–211.PubMedCrossRef
37.
Zurück zum Zitat LeBoeuf M, Terrell A, Trivedi S, Sinha S, Epstein JA, Olson EN, et al. Hdac1 and Hdac2 act redundantly to control p63 and p53 functions in epidermal progenitor cells. Dev Cell. 2010;19(6):807–18.PubMedPubMedCentralCrossRef LeBoeuf M, Terrell A, Trivedi S, Sinha S, Epstein JA, Olson EN, et al. Hdac1 and Hdac2 act redundantly to control p63 and p53 functions in epidermal progenitor cells. Dev Cell. 2010;19(6):807–18.PubMedPubMedCentralCrossRef
38.
Zurück zum Zitat Hughes MW, Jiang TX, Lin SJ, Leung Y, Kobielak K, Widelitz RB, Chuong CM. Disrupted ectodermal organ morphogenesis in mice with a conditional histone deacetylase 1, 2 deletion in the epidermis. J Invest Dermatol. 2014;134(1):24–32.PubMedCrossRef Hughes MW, Jiang TX, Lin SJ, Leung Y, Kobielak K, Widelitz RB, Chuong CM. Disrupted ectodermal organ morphogenesis in mice with a conditional histone deacetylase 1, 2 deletion in the epidermis. J Invest Dermatol. 2014;134(1):24–32.PubMedCrossRef
39.
Zurück zum Zitat Mottet D, Bellahcène A, Pirotte S, Waltregny D, Deroanne C, Lamour V, et al. Histone deacetylase 7 silencing alters endothelial cell migration, a key step in angiogenesis. Circ Res. 2007;101(12):1237–46.PubMedCrossRef Mottet D, Bellahcène A, Pirotte S, Waltregny D, Deroanne C, Lamour V, et al. Histone deacetylase 7 silencing alters endothelial cell migration, a key step in angiogenesis. Circ Res. 2007;101(12):1237–46.PubMedCrossRef
40.
Zurück zum Zitat Urbich C, Rössig L, Kaluza D, Potente M, Boeckel JN, Knau A, et al. HDAC5 is a repressor of angiogenesis and determines the angiogenic gene expression pattern of endothelial cells. Blood. 2009;113(22):5669–79.PubMedCrossRef Urbich C, Rössig L, Kaluza D, Potente M, Boeckel JN, Knau A, et al. HDAC5 is a repressor of angiogenesis and determines the angiogenic gene expression pattern of endothelial cells. Blood. 2009;113(22):5669–79.PubMedCrossRef
41.
Zurück zum Zitat Trivedi CM, Zhu W, Wang Q, Jia C, Kee HJ, Li L, et al. Hopx and Hdac2 interact to modulate Gata4 acetylation and embryonic cardiac myocyte proliferation. Dev Cell. 2010;19(3):450–9.PubMedPubMedCentralCrossRef Trivedi CM, Zhu W, Wang Q, Jia C, Kee HJ, Li L, et al. Hopx and Hdac2 interact to modulate Gata4 acetylation and embryonic cardiac myocyte proliferation. Dev Cell. 2010;19(3):450–9.PubMedPubMedCentralCrossRef
42.
Zurück zum Zitat Karvonen U, Jänne OA, Palvimo JJ. Androgen receptor regulates nuclear trafficking and nuclear domain residency of corepressor HDAC7 in a ligand-dependent fashion. Exp Cell Res. 2006;312(16):3165–83.PubMedCrossRef Karvonen U, Jänne OA, Palvimo JJ. Androgen receptor regulates nuclear trafficking and nuclear domain residency of corepressor HDAC7 in a ligand-dependent fashion. Exp Cell Res. 2006;312(16):3165–83.PubMedCrossRef
43.
Zurück zum Zitat Kwack MH, Shin SH, Kim SR, Im SU, Han IS, Kim MK, et al. l-Ascorbic acid 2-phosphate promotes elongation of hair shafts via the secretion of insulin-like growth factor-1 from dermal papilla cells through phosphatidylinositol 3-kinase. Br J Dermatol. 2009;160(6):1157–62.PubMedCrossRef Kwack MH, Shin SH, Kim SR, Im SU, Han IS, Kim MK, et al. l-Ascorbic acid 2-phosphate promotes elongation of hair shafts via the secretion of insulin-like growth factor-1 from dermal papilla cells through phosphatidylinositol 3-kinase. Br J Dermatol. 2009;160(6):1157–62.PubMedCrossRef
44.
Zurück zum Zitat Chavanas S, Méchin MC, Nachat R, Adoue V, Coudane F, Serre G, Simon M. Peptidylarginine deiminases and deimination in biology and pathology: relevance to skin homeostasis. J Dermatol Sci. 2006;44(2):63–72.PubMedCrossRef Chavanas S, Méchin MC, Nachat R, Adoue V, Coudane F, Serre G, Simon M. Peptidylarginine deiminases and deimination in biology and pathology: relevance to skin homeostasis. J Dermatol Sci. 2006;44(2):63–72.PubMedCrossRef
45.
Zurück zum Zitat Basmanav ÜFB, Cau L, Tafazzoli A, Méchin MC, Wolf S, Romano MT, et al. Mutations in three genes encoding proteins involved in hair shaft formation cause uncombable hair syndrome. Am J Hum Genet. 2016;99(6):1292–304.CrossRef Basmanav ÜFB, Cau L, Tafazzoli A, Méchin MC, Wolf S, Romano MT, et al. Mutations in three genes encoding proteins involved in hair shaft formation cause uncombable hair syndrome. Am J Hum Genet. 2016;99(6):1292–304.CrossRef
46.
Metadaten
Titel
Mapping of cis-acting expression quantitative trait loci in human scalp hair follicles
verfasst von
Marisol Herrera-Rivero
Lara M. Hochfeld
Sugirthan Sivalingam
Markus M. Nöthen
Stefanie Heilmann-Heimbach
Publikationsdatum
01.12.2020
Verlag
BioMed Central
Erschienen in
BMC Dermatology / Ausgabe 1/2020
Elektronische ISSN: 1471-5945
DOI
https://doi.org/10.1186/s12895-020-00113-y

Weitere Artikel der Ausgabe 1/2020

BMC Dermatology 1/2020 Zur Ausgabe

Leitlinien kompakt für die Dermatologie

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Riesenzellarteriitis: 15% der Patienten sind von okkulter Form betroffen

16.05.2024 Riesenzellarteriitis Nachrichten

In einer retrospektiven Untersuchung haben Forschende aus Belgien und den Niederlanden die okkulte Form der Riesenzellarteriitis genauer unter die Lupe genommen. In puncto Therapie und Rezidivraten stellten sie keinen sehr großen Unterschied zu Erkrankten mit kranialen Symptomen fest.

Betalaktam-Allergie: praxisnahes Vorgehen beim Delabeling

16.05.2024 Pädiatrische Allergologie Nachrichten

Die große Mehrheit der vermeintlichen Penicillinallergien sind keine. Da das „Etikett“ Betalaktam-Allergie oft schon in der Kindheit erworben wird, kann ein frühzeitiges Delabeling lebenslange Vorteile bringen. Ein Team von Pädiaterinnen und Pädiatern aus Kanada stellt vor, wie sie dabei vorgehen.

Isotretinoin: Risiko für schwere Laboranomalien „marginal erhöht“

08.05.2024 Akne Nachrichten

Die Aknetherapie mit Isotretinoin kann einen Anstieg von Leberenzymen und Blutfetten verursachen. Das Risiko für schwere Störungen ist laut einer Forschungsgruppe der Universität Lübeck aber nur marginal erhöht und auf einen engen Zeitraum konzentriert.

Darf man die Behandlung eines Neonazis ablehnen?

08.05.2024 Gesellschaft Nachrichten

In einer Leseranfrage in der Zeitschrift Journal of the American Academy of Dermatology möchte ein anonymer Dermatologe bzw. eine anonyme Dermatologin wissen, ob er oder sie einen Patienten behandeln muss, der eine rassistische Tätowierung trägt.

Update Dermatologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.