scholarly journals The Human β Globin Locus Introduced by YAC Transfer Exhibits a Specific and Reproducible Pattern of Developmental Regulation in Transgenic Mice

Blood ◽  
1997 ◽  
Vol 90 (11) ◽  
pp. 4602-4609 ◽  
Author(s):  
Susanna Porcu ◽  
Michael Kitamura ◽  
Ewa Witkowska ◽  
Zemin Zhang ◽  
Annick Mutero ◽  
...  

The human β globin locus spans an 80-kb chromosomal region encompassing both the five expressed globin genes and the cis-acting elements that direct their stage-specific expression during ontogeny. Sequences proximal to the genes and in the locus control region, 60 kb upstream of the adult β globin gene, are required for developmental regulation. Transgenic studies have shown that altering the structural organization of the locus disrupts the normal pattern of globin gene regulation. Procedures for introducing yeast artificial chromosomes (YACs) containing large genetic loci now make it possible to define the sequences required for stage-restricted gene expression in constructs that preserve the integrity of the β globin locus. We demonstrate that independent YAC transgenic lines exhibit remarkably similar patterns of globin gene expression during development. The switch from γ to β globin predominant expression occurs between day 11.5 and 12.5 of gestation, with no more than twofold differences in human β globin mRNA levels between lines. Human β globin mRNA levels were twofold to fourfold lower than that of mouse βmaj, revealing potentially significant differences in the regulatory sequences of the two loci. These findings provide an important basis for studying regulatory elements within the β globin locus.

Blood ◽  
1997 ◽  
Vol 90 (11) ◽  
pp. 4602-4609 ◽  
Author(s):  
Susanna Porcu ◽  
Michael Kitamura ◽  
Ewa Witkowska ◽  
Zemin Zhang ◽  
Annick Mutero ◽  
...  

Abstract The human β globin locus spans an 80-kb chromosomal region encompassing both the five expressed globin genes and the cis-acting elements that direct their stage-specific expression during ontogeny. Sequences proximal to the genes and in the locus control region, 60 kb upstream of the adult β globin gene, are required for developmental regulation. Transgenic studies have shown that altering the structural organization of the locus disrupts the normal pattern of globin gene regulation. Procedures for introducing yeast artificial chromosomes (YACs) containing large genetic loci now make it possible to define the sequences required for stage-restricted gene expression in constructs that preserve the integrity of the β globin locus. We demonstrate that independent YAC transgenic lines exhibit remarkably similar patterns of globin gene expression during development. The switch from γ to β globin predominant expression occurs between day 11.5 and 12.5 of gestation, with no more than twofold differences in human β globin mRNA levels between lines. Human β globin mRNA levels were twofold to fourfold lower than that of mouse βmaj, revealing potentially significant differences in the regulatory sequences of the two loci. These findings provide an important basis for studying regulatory elements within the β globin locus.


Blood ◽  
2006 ◽  
Vol 108 (11) ◽  
pp. 555-555 ◽  
Author(s):  
Hassana Fathallah ◽  
Ali Taher ◽  
Ali Bazarbachi ◽  
George F. Atweh

Abstract A number of therapeutic agents including hydroxyurea, butyrate and decitabine have shown considerable promise in the treatment of sickle cell disease (SCD). However, the same agents have shown less clinical activity in β-thalassemia. As a first step towards understanding the molecular basis of the different clinical responses to these agents, we have studied the mechanisms of induction of fetal hemoglobin (HbF) by butyrate in BFU-E derived cells from 5 patients with SCD and 9 patients with β-thalassemia intermedia. Exposure to butyrate resulted in a dose-dependent augmentation of γ-globin mRNA levels in erythroid cells from patients with SCD. In contrast, induction of γ-globin expression in erythroid cells from patients with β-thalassemia intermedia was only seen at a high concentration of butyrate. The increase in γ-globin mRNA levels in patients with SCD and β-thalassemia intermedia was associated with opening of the DNA structure as manifested by decreased DNA methylation at the γ-globin promoters. Interestingly, butyrate exposure had markedly different effects on the expression of the β- and α-globin genes in the two categories of patients. Butyrate decreased the level of β-globin mRNA in 4 out of 5 patients with SCD (P = 0.04), while in β-thalassemia the levels of β-globin mRNA did not change in 7 patients and decreased in 2 patients after butyrate exposure (P = 0.12). Thus in patients with SCD, the effects of the induction of the γ-globin gene on the γ/(β+γ) mRNA ratios were further enhanced by the butyrate-mediated decreased expression of the β-globin gene. As a result, γ/(β+γ) mRNA ratios increased in all patients with SCD, with a mean increase of 31% (P = 0.002). In contrast, butyrate increased γ/(β+γ) mRNA ratios only in 4 out of 9 patients with β-thalassemia, with a more modest mean increase of 12% (P = 0.004). Interestingly, the decreased β-globin expression in patients with SCD was associated with closing of the DNA configuration as manifested by hypermethylation of DNA at the promoter of the β-globin gene while methylation of the same promoter did not change following butyrate exposure in patients with β-thalassemia intermedia. More surprisingly, the expression of the α-globin genes increased following butyrate exposure in 4 out of 9 patients with β-thalassemia, while the levels of α-globin mRNA decreased in 4 out of 5 patients with SCD. As a result, the favorable effects of the butyrate-induced increase in γ-globin gene expression on the α: non-α mRNA imbalance in patients with β-thalassemia intermedia were partly neutralized by the corresponding increase in α-globin gene expression. These differences may explain, at least in part, the more favorable effects of inducers of HbF in SCD than in β-thalassemia. Further studies are necessary to fully understand the molecular bases of the different responses to agents that induce HbF in patients with these disorders.


1994 ◽  
Vol 72 (9-10) ◽  
pp. 377-380 ◽  
Author(s):  
Christi Andrin ◽  
Charlotte Spencer

Gene expression is an extremely complicated process in which several mechanisms are involved. Owing to its developmental and tissue-specific expression, the β-globin gene is an excellent model for studying gene expression. β-Globin gene expression involves an interplay between several different mechanisms. Chromatin structure is thought to be altered by the locus control region (LCR) located far upstream of the β-globin gene locus. As well, multiple transcription factors come into play both in the LCR and in the individual promoters and enhancers of the β-globin genes. The interaction between these then allows for delicate regulation of β-globin gene expression. In the following review the elaborate system of β-globin gene expression will briefly be examined.Key words: β-globin, gene expression, chromatin, GATA-I, NF-E2, developmental regulation.


1987 ◽  
Vol 7 (1) ◽  
pp. 398-402
Author(s):  
T Rutherford ◽  
A W Nienhuis

The contribution of the human globin gene promoters to tissue-specific transcription was studied by using globin promoters to transcribe the neo (G418 resistance) gene. After transfection into different cell types, neo gene expression was assayed by scoring colony formation in the presence of G418. In K562 human erythroleukemia cells, which express fetal and embryonic globin genes but not the adult beta-globin gene, the neo gene was expressed strongly from a fetal gamma- or embryonic zeta-globin gene promoter but only weakly from the beta promoter. In murine erythroleukemia cells which express the endogenous mouse beta genes, the neo gene was strongly expressed from both beta and gamma promoters. In two nonerythroid cell lines, human HeLa cells and mouse 3T3 fibroblasts, the globin gene promoters did not allow neo gene expression. Globin-neo genes were integrated in the erythroleukemia cell genomes mostly as a single copy per cell and were transcribed from the appropriate globin gene cap site. We conclude that globin gene promoter sequences extending from -373 to +48 base pairs (bp) (relative to the cap site) for the beta gene, -385 to +34 bp for the gamma gene, and -555 to +38 bp for the zeta gene are sufficient for tissue-specific and perhaps developmentally specific transcription.


Blood ◽  
2005 ◽  
Vol 106 (11) ◽  
pp. 523-523
Author(s):  
Marco De Gobbi ◽  
Vip Viprakasit ◽  
Pieter J. de Jong ◽  
Yuko Yoshinaga ◽  
Jan-Fang Cheng ◽  
...  

Abstract The human α globin cluster includes an embryonic gene ζ and 2 fetal/adult genes (α2 and α1) arranged along the chromosome in the order in which they are expressed in development (5′-ζ-pseudoζ- αD- α2-α1-𝛉-3′). Fully activated expression of these genes in erythroid cells depends on upstream regulatory elements of which HS-40, located 40kb upstream of the cluster, appears to exert the greatest effect. We have recently shown that during terminal differentiation, key transcription factors (GATA-2, GATA-1, NF-E2, SCL complex) sequentially bind the α promoters and their regulatory elements and a domain of histone acetylation develops which eventually encompasses the entire α globin cluster including the upstream regulatory sequences. α-thalassemia most frequently results from deletions or point mutations affecting the structural α globin genes, but may also result from rare sporadic deletions which remove the upstream regulatory sequences. In a single family α globin expression was silenced by a mutation which drives an anti-sense RNA through the α gene. Alpha thalassemia may also result from inherited and acquired mutations in a trans-acting factor called ATRX. Over the past few years we have continued to screen for new mechanisms which lead to α thalassemia and thereby elucidate new principles underlying the regulation of gene expression in hemopoiesis. Here we describe a new mechanism of α thalassemia occurring in Pacific Islanders in whom we could detect no mutations or rearrangements in the α globin gene locus. Despite this, extensive genetic analysis showed unequivocally that the causative mutation is linked to the terminal 169kb of chromosome 16 (Viprakasit et al accompanying abstract). Analysis of globin synthesis, steady state RNA levels and detection of RNA in situ demonstrated that the mutation downregulates α globin transcription. To identify the mutation, we constructed a new BAC library from an affected homozygote, isolated and re-sequenced the candidate region and focussed further analysis on 8 SNPS within the α globin cluster, one of which creates a new GATA-1 binding site (GACA>GATA). Using primary erythroblasts from normal individuals and patients with this form of thalassemia, together with interspecific hybrids containing either the normal or abnormal copy of chromosome 16, we have shown that this SNP creates a new binding site in vivo for GATA-1 and the SCL complex. Furthermore, the chromatin at this site becomes activated as judged by acetylation of histone H3 and H4 (H3ac2 and H4ac4) and methylation of histone H3 (H3K4me2). Based on these data we postulate that an active transcriptional complex binding this new GATA site created by the SNP-mutation, could distract the upstream regulatory regions, which normally interact with the α globin promoter, and silence α globin gene expression. This model thus represents a new example of α globin gene down-regulation and a new mechanism by which gene expression can be perturbed during hemopoiesis.


Blood ◽  
2006 ◽  
Vol 108 (11) ◽  
pp. 1581-1581
Author(s):  
Rodwell Mabaera ◽  
Christine Richardson ◽  
Sarah Conine ◽  
Christopher H. Lowrey

Abstract 5-Azacytidine (5-Aza) was demonstrated to be a potent inducer of human fetal globin gene expression more than 20 years ago. More recently, 5-Aza-2-deoxycytidine has been shown to have similar properties. Since the 1980’s there have been two predominant hypotheses to explain the action of these agents. The first is based on the observation that these, and several other active inducing agents, are cytotoxic to differentiating erythroid cells and that drug treatment alters the kinetics of erythroid differentiation. This has been proposed to result in prolonged expression of the γ-globin genes which are normally expressed only early in differentiation. The second is based on the observation that both agents are DNA methyltransferase inhibitors and are presumed to cause demethylation of cellular DNA including the γ-globin gene promoters leading to activation of the genes. These two models lead to specific predictions that we have evaluated using an in vitro erythroid differentiation system. In this system, human adult CD34+ cells are cultured in SCF, Flt3 ligand and IL-3 for 7 days and then switched to Epo for 14 days. This results in an exponential expansion of erythroid cells. As has been described for normal human differentiation, these cells express small amounts of γ-globin mRNA early in differentiation followed by a much larger amount of β-globin mRNA. HPLC at the end of the culture period shows 99% HbA and 1% HbF. Treatment of cultures on a daily basis with 5-Aza starting on day 10 results in dose dependent increases in γ-globin mRNA, Gγ- and Aγ-chain production and HbF. The cytotoxicity model predicts that γ-globin expression will be prolonged to later in differentiation - and this is seen. However, a daily 5-Aza dose of 300 nM, which produces ~80% of the maximal response in γ-globin mRNA and HbF, has no effect on cell growth or differentiation kinetics. This argues against the toxicity model. We next examined the effect of 5-Aza on γ-globin promoter methylation using the bisulfite method. We studied CpGs at −344, −252, −162, −53, −50, +6, +19 and +50 relative to the start site. For untreated controls, all of the sites are nearly 100% methylated at day 1. By day 3, the upstream sites become ~50% methylated except the −53 CpG which was <20%. This pattern persisted at day 10. By day 14 the promoters had become largely remethylated. For cells treated with 5-Aza starting on day 10, there was no change in the levels of methylation seen on days 1,3 and 10, but at day 14 the low levels of upstream methylation persisted - just as γ-globin expression does. However, in both treated and untreated cells, down-stream CpG sites were highly methylated at all time points. This suggests that γ promoter demethylation may be due to a local and not a generalized effect of 5-Aza on cellular DNA methylation. We also made two unexpected observations. At a 300nM dose of 5-Aza, γ-globin mRNA is ~doubled while β-globin mRNA levels are ~halved - indicating that 5-Aza not only induces γ-globin expression also suppresses β-globin. Also despite only a doubling in γ-globin mRNA, there was an ~50-fold increase in HbF, from ~1% to more than 50%, while total per cell Hb levels were unchanged. Neither of these results are easily explained by current models of γ-globin gene induction. Our results raise the possibility that mechanisms beyond cytotoxicity and generalized DNA demethylation may be responsible for pharmacologic induction of γ-globin mRNA and HbF.


Blood ◽  
2005 ◽  
Vol 106 (11) ◽  
pp. 820-820
Author(s):  
Wei Li ◽  
Betty S. Pace

Abstract The design and evaluation of therapies for sickle cell disease (SCD) rely on our understanding of hemoglobin accumulation during erythropoiesis and sequential globin gene expression (ε → Gγ → Aγ → δ → β) during development. To gain insights into globin gene switching, we completed time course micorarray analyses of erythroid progenitors to identify trans-factors involved in γ gene activation. Studies were completed to map the pattern of γ and β globin gene expression in progenitors grown from normal peripheral blood mononuclear cells. We compared cells grown in a 2-phase (phase 1, d0-6: SCF, IL-3, IL-6, and GM-CSF and phase 2, d7-25: SCF and EPO) vs. 1-phase (d0-34: SCF, IL-3, and EPO) liquid culture system. From day 0 to 34 in either system cell viability remained &gt;99%. Total RNA was isolated using Trizol and column cleanup (Qiagen). Globin mRNA levels were measured at 2–3 day intervals by quantitative PCR (qPCR). In the 2-phase system γ-globin mRNA&gt;β-globin mRNA up to d14, 4 days of approximately equal expression then β mRNA &gt; γ mRNA by d20. By contrast, in 1-phase studies there was a rapid switch around d20(see graph). We speculate that this difference may be due to the early addition of EPO on d0 therefore we continued our detailed analysis in this system. To confirm that our in vitro system recapitulates in vivo gene expression patterns, we completed studies to ascertain Gγ - vs. Aγ globin mRNA levels. The normalized Gγ:Aγ ratio decreased from ~3:1 on d7 to ~1:1 by d34; These findings were confirmed using two sets of Gγ and Aγ globin primers. We concluded that the 1-phase system recapitulated normal γ/β globin switching and that gene profiling studies to identify the trans-factor involved in switching mechanisms were feasible. We used Discover oligo chips (ArrayIt, Sunnyvale, CA) containing 380 human genes selected from 30 major functional groups including hematopoiesis. To aide interpretation of chip data, cell populations were rated morphologically using Giemsa stained cytospin preps. From d16 on we observed an increase in late erythroid progenitors (normoblasts) from 1% to 71% by d31. After verifying RNA quality by gel inspection of ribosomal molecules, we prepared Cy3 and Cy5 probes for early and late time-point RNA samples respectively. Chip analysis was performed at several time points but d0/21, d7/21, and d21/28 were most informative. Based on Axon GenePixPro 6.0 and Acuity 4.0 software analysis we found the following genes with &gt;1.5-fold change in expression profile (shown as down-regulated/up-regulated genes): d0/21: 33/73, d7/21: 13/25, and d21/28:35/26. Principal component analysis (PCA), hierarchical clusters and self organizing maps were constructed. Gene profiles were correlated with the γ/β switching curve using d7 (γ &gt;β), d21 (γ ~ β), and d28 (γ &lt;β) data. Hematopoietic dataset analysis at d21 revealed 4 candidate γ-globin gene activators including v-myb, upsteam binding transfactor -RNApol1 and 2 zinc finger proteins. Analysis of a d28 dataset revealed 12 proteins involved in γ-globin gene silencing including IL-3, SCF, MAPKKK3, v-raf-1, ATF-2, and glucocorticoid receptor DNA binding factor 1 among others. Gene expression profiles will be validated using qPCR and promising candidates will be tested by forced expression in transient and stable reporter systems. Figure Figure


Blood ◽  
2007 ◽  
Vol 110 (11) ◽  
pp. 4102-4102
Author(s):  
Vladan P. Cokic ◽  
Bojana B. Beleslin-Cokic ◽  
Constance Tom Noguchi ◽  
Alan N. Schechter

Abstract We have previously shown that nitric oxide (NO) is involved in the hydroxyurea-induced increase of gamma-globin gene expression in cultured human erythroid progenitor cells and that hydroxyurea increases NO production in endothelial cells via endothelial NO synthase (NOS). Here we report that co-culture of human bone marrow endothelial cells with erythroid progenitor cells induced gamma-globin mRNA expression (1.8 fold), and was further elevated (2.4 fold) in the presence of hydroxyurea (40 μM). Based on these results, NOS-dependent stimulation of NO levels by bradykinin and lipopolysaccharide has been observed in endothelial (up to 0.3 μM of NO) and macrophage cells (up to 6 μM of NO), respectively. Bradykinin slightly increased gamma-globin mRNA levels in erythroid progenitor cells, but failed to increase gamma-globin mRNA levels in endothelial/erythroid cell co-cultures indicating that stimulation of endothelial cell production of NO alone is not sufficient to induce gamma-globin expression. In contrast, lipopolysaccharide and interferon-gamma mutually increased gamma-globin gene expression (2 fold) in macrophage/erythroid cell co-cultures. In addition, hydroxyurea (5–100 μM) induced NOS-dependent production of NO in human (up to 0.7 μM) and mouse macrophages (up to 1.2 μM). Co-culture studies of macrophages with erythroid progenitor cells also resulted in induction of gamma-globin mRNA expression (up to 3 fold) in the presence of hydroxyurea (20–100 μM). These results demonstrate a mechanism by which hydroxyurea may induce globin genes and affect changes in the phenotype of hematopoietic cells via the common paracrine effect of bone marrow stromal cells.


Blood ◽  
1998 ◽  
Vol 92 (9) ◽  
pp. 3416-3421 ◽  
Author(s):  
E. Skarpidi ◽  
G. Vassilopoulos ◽  
G. Stamatoyannopoulos ◽  
Q. Li

To examine whether transfer of γ globin genes into mouse erythroleukemia cells can be used for the analysis of regulatory elements of γ globin gene promoter, Aγ gene constructs carrying promoter truncations that have been previously analyzed in transgenic mice were used for production of stably transfected mouse erythroleukemia (MEL) cell clones and pools. We found that constructs, which contain a microlocus control region (μLCR) that efficiently protects globin gene expression from the effects of the position of integration in transgenic mice, display position-dependent globin gene expression in MEL cell clones. Aγ globin gene expression among MEL cell clones carrying the μLCR(−201)Aγ and μLCR(−382)Aγ gene constructs ranged 15.5-fold and 17.6-fold, respectively, and there was no correlation between theAγ mRNA levels and the copies of the transgene (r= .28, P = .18). There was significant variation in per copy Aγ globin gene expression among MEL cell pools composed of 10 clones, but not among pools composed of 50 clones, indicating that position effects are averaged in pools composed by large numbers of clones. The overall pattern of Aγ globin gene expression in MEL cell pools resembled that observed in transgenic mice indicating that MEL cell transfections can be used in the study ofcis elements controlling γ globin gene expression. MEL cell transfections, however, are not appropriate for investigation of cis elements, which either sensitize or protect the globin transgenes from position effects. © 1998 by The American Society of Hematology.


Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 54-54 ◽  
Author(s):  
Heather L Edward ◽  
Tasha Morrison ◽  
Jacqueline N Milton ◽  
Hong-yuan Luo ◽  
Lance Davis ◽  
...  

Abstract Hereditary persistence of fetal hemoglobin (HPFH) and (δβ)0 thalassemia are caused by deletions within the β-globin gene (HBB) cluster that remove elements that affect the expression of the γ-globin genes (HBG2 and HBG1, or HBG). These deletions are of different lengths and have different 5’ and 3’ breakpoints. The phenotypes associated with heterozygous carriers of (δβ)0 thalassemia and HPFH deletions are differentiated by levels of 5-15% HbF distributed heterocellularly in the former and 15-30% HbF distributed pancellularly in the latter. We found a novel 588.6 kb deletion that removed both the 3.5 kb fragment 5’ to HBD that is deleted in Corfu β thalassemia and contains a BCL11A binding site, and the known cis-acting elements downstream of HBB. The proband with this deletion had a HbF of 5.4% (Morrison et al, Blood, 2014 abstract 3452). To study the relative importance of 5’ and 3’ regulatory elements in HBG expression we studied 209 cases culled from the literature and from our laboratory where the 3.5 kb element 5’ to HBD and enhancers 3’ to HBB were deleted and HBG remained intact. We used a backwards stepwise regression statistical analysis to determine which deleted elements had the greatest effect on HbF levels. The combination of the deletion of 3.5 kb intergenic region 5’ to HBD, the presence of the HPFH-1 “3D” enhancer juxtaposed to HBG, and the deletion of the 3’ HS1 region accounted for 66.7% of the HbF variation in heterozygotes for HPFH and (δβ)0-thalassemia deletions. The HPFH-1 “3D” enhancer juxtaposed to HBG— the main difference between HPFH-1 and 2 compared with Spanish (δβ)0-thalassemia—was associated with an increase in HbF of 20.78% (p<2e-16) after adjusting for the effects of the other 5’ and 3’ cis-acting elements. The next most significant factor was the deletion of the 3.5 kb fragment 5’ to HBD which resulted in an increase of 10.62% HbF after similar adjustments (p<2e-16); deletion of the 3’ HS1 region accounted for an increase in HbF of 5.25% (p<1.05e-5). The HPFH-3 and HPFH-6 enhancer regions each accounted for a less than 1% increase in HbF and were not significantly associated with HbF in this model. Among 194 individuals where both 5’ and some 3’ elements affecting γ-globin gene expression—excluding the “3D” enhancer—were deleted, HbF was 20±9.3%; in 13 cases where all 3’ enhancers—including the “3D” enhancer—were deleted, HbF was 6.8±3.7% (p=8.9e-07). To determine which combinations of cis-acting elements were associated with high and low HbF levels we performed a classification and regression tree (cART) analysis on HbF. The results of the regression tree (Figure) only included the deletion of the 5’ 3.5 kb fragment region, the presence of the HPFH-1 “3D” enhancer and the deletion of the 3’ HS1 region and were consistent with the results of the backwards selection model. The absence of the 5’ 3.5 kb fragment 5’ to HBD combined with the presence of the HPFH-1 “3D” enhancer was associated with the highest average HbF of 27.02%. The absence of the 3.5 kb fragment 5’ to HBD combined with the absence of the HPFH-1 “3D” enhancer was associated with the lowest average HbF of 6.82%.The 588.6 kb deletion is the largest deletion reported in the HBB cluster that leaves the γ-globin genes intact, and the second to remove both the BCL11A binding site and all known 3’ enhancer elements. By studying deletions in the HBBgene cluster we have further defined the hierarchy of cis-acting elements that modulate HbF levels in adults and suggest a paramount role of the distal “3D” enhancer. Figure 1 Figure 1. Disclosures No relevant conflicts of interest to declare.


Sign in / Sign up

Export Citation Format

Share Document