p38 MAPK Signalling Is Not Sufficient to Induce γ-Globin Gene Expression

Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 1083-1083
Author(s):  
Emily K Schaeffer ◽  
Rachel J West ◽  
Cynthia K Hahn ◽  
Christopher H. Lowrey

Abstract Abstract 1083 Several recent publications have implicated p38 MAPK signalling in the pharmacologic induction of γ-globin gene expression. These conclusions are primarily based on demonstrations of increased p38 phosphorylation (P-p38) by inducing compounds and inhibition of induction by the p38 inhibitor SB203580 (SB). Based on these studies and work from our own lab, we proposed that p38 signalling might be the key mechanism underlying γ-globin gene induction by most active agents. We have performed a series of experiments in K562 cells designed to critically test this hypothesis and to better define the role of p38 signalling in γ-globin gene induction. We first performed dose-response experiments with sodium butyrate (NaB), an agent that induces γ-globin mRNA in vitro and in vivo and has been proposed to work through p38 signalling. If p38 signalling is inducing γ-globin gene expression, then p38 phosphorylation should correlate with increased γ-globin mRNA levels. We found 3–3.5-fold induction of steady-state γ-globin mRNA at doses ranging from 0.25 to 2mM NaB. SB suppressed induction at all dose levels. However, P-p38 was increased only at doses below 1mM and was actually suppressed compared to untreated controls at 2mM NaB. We next examined the effects of heat shock and UV and X-irradiation, all well-known activators of p38 signalling. Here too we found specific stress “doses” that did not increase P-p38 but still strongly induced γ-globin mRNA and were inhibited by SB. These results demonstrated a lack of correlation between p38 phosphorylation and γ-globin induction but for every experimental condition tested, SB inhibits basal or induced γ-globin mRNA levels. To more closely examine this discrepancy, we transiently expressed a constitutively active MKK6 mutant (MKK6-Glu; the direct upstream activator of p38), to mimic transient increases in P-p38 seen with drugs and stresses. While MKK6-Glu expression elevated p38 phosphorylation to levels seen with NaB and physical stresses, it did not increase basal γ-globin mRNA levels. Furthermore, increased p38 activation by MKK6-Glu did not enhance NaB-induced γ-globin expression. We next used p38 siRNA in combination with NaB and heat shock conditions that induce γ-globin mRNA. This allowed us to decrease downstream p38 signalling (as judged by Hsp27 phosphorylation) to levels equivalent to those achieved with SB. While SB still completely inhibited induction, there was only a partial decrease in NaB induction and no effect on heat shock induction with the p38 knockdown. Our results show that p38 phosphorylation does not strictly correlate with γ-globin induction suggesting that it is not required for at least some degree of γ-globin induction by NaB and stress. The fact that we saw no γ-globin mRNA induction with targeted activation of p38 by MKK6-Glu suggests that p38 activation is not sufficient for induction. Our results also suggest that SB inhibits γ-globin mRNA induction by a mechanism other than direct p38 inhibition. Upon investigating potential mechanisms, we observed SB to promote ERK phosphorylation, suggesting that a modulation of ERK signalling or p38/ERK cross-talk may be involved. Additionally, we find that SB treatment appears to decrease steady-state mRNA levels but not nascent γ-globin mRNA, suggesting a role in transcript processing or stability. Disclosures: No relevant conflicts of interest to declare.

Blood ◽  
2009 ◽  
Vol 114 (22) ◽  
pp. 459-459
Author(s):  
Rachel West ◽  
Rodwell Mabaera ◽  
Sarah Conine ◽  
Elizabeth Macari ◽  
William J. Lowrey ◽  
...  

Abstract Abstract 459 We have recently presented a cell stress signaling model to explain the ability of a wide variety of mechanistically distinct drugs to induce fetal hemoglobin (HbF) production (Mabaera et al, Exp Hematol 2008). This model proposes that HbF inducing agents, including DNMT and HDAC inhibitors, short-chain fatty acid derivatives and cytotoxic agents induce HbF production by activating intracellular cell stress signaling pathways and that the p38 MAPK (mitogen-activated protein kinase) pathway plays a central role in the induction process. The major alternative to this model is that each class of drugs works through a distinct mechanism, such as global DNA hypomethylation or histone hyperacetylation. If our model is correct, then non-pharmacologic inducers of p38 MAPK signaling, such as X-ray and UV irradiation, heat shock, and osmotic shock should also increase γ-globin gene expression. If these physical stresses do not up-regulate γ-globin gene expression, then our model is likely to be incorrect. To mimic our previous experiments in which we treated differentiating human primary erythroid cells with inducing agents over several days, we exposed K562 cells to physical stresses daily for 5 days at doses that did not cause cell death. To test whether X-irradiation induces the transcription of γ-globin mRNA, cells were irradiated at doses ranging from 0 to 1.0 cGy/day, in the absence and presence of SB203580 (a p38 MAPK inhibitor). Experiments were first performed on K562 cells, and then during in vitro erythroid differentiation of primary human CD34+ cells. These treatments stimulated a strong, dose-dependent increase of γ-globin mRNA in both K562 cells (up to 6-fold over untreated cells) and in primary erythroid cells (up to 5-fold) as measured by quantitative RT-PCR. SB203580 abolished this effect in both cell types. The inhibitor also caused a decrease in γ-globin mRNA expression in untreated control cells, suggesting that p38 MAPK signaling plays a role in basal γ-globin gene expression. The same X-ray doses also induced phosphorylation of the down-stream p38 MAPK target, HSP-27 and up-regulated expression of stress-induced transcription factor genes including GADD34, CHOP (GADD153), and ATF3. Expression of GADD34 and CHOP genes and HSP-27 phosphorylation were inhibited by SB203580. We next tested the ability of UV light (254 nm) administered daily for 5 days to induce γ-globin mRNA in K562 cells. A dose-dependent increase in γ-globin mRNA was observed following exposure to doses as low as 10 J/m2. Cells treated with 35 J/m2 had 2.7-fold higher levels of steady state γ-globin mRNA compared to untreated control cells. This induction was also inhibited by SB203580. Preliminary experiments with heat shock have yielded similar results. Following a single 180-minute exposure to 42°C, K562 cells showed a 2.2-fold increase in γ-globin mRNA compared to untreated cells. Shorter exposures to higher temperatures (e.g., 50°C for 15 minutes) caused an approximately 3-fold increase in γ-globin steady-state mRNA after 24 hours. These 2- to 3-fold increases in expression are similar to those we have previously observed in primary human erythroid cells with 5-azacytidine and butyrate. In none of our experiments did the K562 cells become benzidine positive, indicating that increased γ-globin expression was not the result of activating an erythroid differentiation program. Together with other published studies, these data support the hypothesis that p38 MAPK pathway signaling, whether caused by drugs or physical stress, is a key component of γ-globin gene induction. This in turn suggests that the components of the p38 MAPK pathway could serve as novel targets for the development of new HbF inducing agents. Disclosures: No relevant conflicts of interest to declare.


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 ◽  
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.


Blood ◽  
2010 ◽  
Vol 116 (21) ◽  
pp. 2074-2074
Author(s):  
Elizabeth R. Macari ◽  
Rachel J. West ◽  
Christopher H. Lowrey

Abstract Abstract 2074 Induction of fetal hemoglobin (HbF) has proven therapeutic potential to treat sickle cell disease and β-thalassemia. However, agents known to be effective in humans, including hydroxyurea, DNMT inhibitors and butyrate derivatives are not ideal due to suppression of hematopoiesis and the possibility of long-term side effects. Two natural compounds, angelicin (Lampronti, et al, Eur J Hematol 2003) and resveratrol (Rodrigue, et al, Hematology 2001) have been found to induce γ-globin gene expression in K562 cells. These agents may be important lead compounds as they are generally non-cytotoxic and are being evaluated in ongoing human trials as cancer chemopreventative agents where these and several other agents are thought to work by activating antioxidant response pathway genes. The products of these genes are enzymes involved in antioxidant and detoxification activities and include NADPH-quinone oxireductase 1 (NQO1), glutamate-cysteine ligase (GCL) and glutathione S-transferase (GST). The activation of these genes is mediated by the transcription factor, NF-E2 related factor 2 (NRF2), which binds to a specific antioxidant response element (ARE) sequence (TGACnnnGCA) in target gene promoters. The proximal γ-globin promoter contains an ARE sequence between the two CAAT boxes, suggesting that it too may be activated by NRF2. This led us to hypothesize that drugs that activate the ARE/NRF2 pathway may provide a less toxic approach to HbF induction. To test this hypothesis, we treated K562 cells with various NRF2 pathway activators. We initially tested six compounds that are known to induce antioxidant response genes at doses that did not inhibit proliferation and found the most pronounced γ-globin induction with tert-butylhydroquinone (tBHQ) (2.8 fold). We next tested tBHQ in two different primary cell culture models: erythroid precursors isolated from normal human bone marrow and in vitro erythroid differentiation of primary human CD34+ cells. In both of these models, tBHQ treatment increased γ-globin steady state mRNA levels and induced expression of NRF2 target genes. Treatment of differentiating erythroid cells caused a dose dependent increase in γ/(γ+β) mRNA and % HbF. HPLC analysis revealed the highest non-toxic concentration of tBHQ, 5μM, produced 10% HbF while the positive control, 0.5μM of 5-Azacytidine, resulted in 12% HbF compared to the untreated control at 3%. Since similar mRNA induction was seen in K562 cells, we used these cells to characterize the mechanism of tBHQ induced γ-globin expression. First, we used siRNA to decrease NRF2 mRNA levels. This resulted in a greater than 75% knockdown of NRF2 mRNA and protein and reduced tBHQ induction of γ-globin and NQO1 gene expression by 90% and 75%, respectively, compared to samples transfected with scrambled siRNA (p < 0.01). Subsequent experiments showed that tBHQ treatment resulted in NRF2 translocation to the nucleus and binding to the NQO1 and γ-globin promoters but not at negative control sites. In addition, inducing NRF2 translocation by transiently suppressing levels of its inhibitor, KEAP1, did not result in full induction of γ-globin mRNA expression, suggesting that NRF2 translocation alone is not sufficient for γ-globin induction. However, when combined with tBHQ, suppression of KEAP1 did enhance γ-globin induction. Taken together, these results suggest that NRF2/ARE pathway activation induces γ-globin mRNA expression and HbF production in primary human erythroid cells and that this is a promising strategy for further pre-clinical development. Disclosures: No relevant conflicts of interest to declare.


Blood ◽  
2019 ◽  
Vol 134 (Supplement_1) ◽  
pp. 812-812
Author(s):  
Mudit Chaand ◽  
Chris Fiore ◽  
Brian T Johnston ◽  
Diane H Moon ◽  
John P Carulli ◽  
...  

Human beta-like globin gene expression is developmentally regulated. Erythroblasts (EBs) derived from fetal tissues, such as umbilical cord blood (CB), primarily express gamma globin mRNA (HBG) and HbF, while EBs derived from adult tissues, such as bone marrow (BM), predominantly express beta globin mRNA (HBB) and adult hemoglobin. Human genetics has validated de-repression of HBG in adult EBs as a powerful therapeutic paradigm in diseases involving defective HBB, such as sickle cell anemia. To identify novel factors involved in the switch from HBG to HBB expression, and to better understand the global regulatory networks driving the fetal and adult cell states, we performed transcriptome profiling (RNA-seq) and chromatin accessibility profiling (ATAC-seq) on sorted EB cell populations from CB or BM. This approach improves upon previous studies that used unsorted cells (Huang J, Dev Cell 2016) or that did not measure chromatin accessibility (Yan H, Am J Hematol 2018). CD34+ cells from CB and BM were differentiated using a 3-phase in vitro culture system (Giarratana M, Blood 2011). Fluorescence-activated cell sorting and the cell surface markers CD36 and GYPA were used to isolate 7 discrete populations, with each sorting gate representing increasingly mature, stage-matched EBs from CB or BM (Fig 1A, B). RNA-seq analysis revealed expected expression patterns of the beta-like globins, with total levels increasing during erythroid maturation and primarily composed of HBB or HBG transcripts in BM or CB, respectively (Fig 1C). Erythroid maturation led to progressive increases in chromatin accessibility at the HBB promoter in BM populations. In CB-derived cells, erythroid maturation led to progressive increases in chromatin accessibility at the HBG promoters through the CD36+GYPA+ stage (Pops 1-5). Chromatin accessibility shifted from the HBG promoters to the HBB promoter during the final stages of differentiation (Pops 6-7), suggesting that HBG gene activation is transient in CB EBs (Fig 1D). Hierarchical clustering and principal component analysis of ATAC-seq data revealed that cell populations cluster based on differentiation stage rather than by BM or CB lineage, suggesting most molecular changes are stage-specific, not lineage-specific (Fig 2A, B). To identify transcription factors driving cell state, and potentially beta-like globin expression preference, we searched for DNA binding motifs within regions of differential chromatin accessibility and found NFI factor motifs enriched under peaks that were larger in BM relative to CB (Fig 2C). Transcription factor footprinting analysis showed that both flanking accessibility and footprint depth at NFI motifs were also increased in BM relative to CB (Fig 2D). Increased chromatin accessibility was observed at the NFIX promoter in BM relative to CB populations, and in HUDEP-2 relative to HUDEP-1 cell lines (Fig 2E). Furthermore, accessibility at the NFIX promoter correlated with elevated NFIX mRNA in BM and HUDEP-2 relative to CB and HUDEP-1, respectively. Together these data implicated NFIX in HbF repression, a finding consistent with previous genome-wide association and DNA methylation studies that suggested a possible role for NFIX in regulating beta-like globin gene expression (Fabrice D, Nat Genet 2016; Lessard S, Genome Med 2015). To directly test the hypothesis that NFIX represses HbF, short hairpin RNAs were used to knockdown (KD) NFIX in primary erythroblasts derived from human CD34+ BM cells (Fig 3A). NFIX KD led to a time-dependent induction of HBG mRNA, HbF, and F-cells comparable to KD of the known HbF repressor BCL11A (Fig 3B-D). A similar effect on HbF was observed in HUDEP-2 cells following NFIX KD (Fig 3E). Consistent with HbF induction, NFIX KD also increased chromatin accessibility and decreased DNA methylation at the HBG promoters in primary EBs (Fig 3F, G). NFIX KD led to a delay in erythroid differentiation as measured by CD36 and GYPA expression (Fig 3H). Despite this delay, by day 14 a high proportion of fully enucleated erythroblasts was observed, suggesting NFIX KD cells are capable of terminal differentiation (Fig 3H). Collectively, these data have enabled identification and validation of NFIX as a novel repressor of HbF, a finding that enhances the understanding of beta-like globin gene regulation and has potential implications in the development of therapeutics for sickle cell disease. Disclosures Chaand: Syros Pharmaceuticals: Employment, Equity Ownership. Fiore:Syros Pharmaceuticals: Employment, Equity Ownership. Johnston:Syros Pharmaceuticals: Employment, Equity Ownership. Moon:Syros Pharmaceuticals: Employment, Equity Ownership. Carulli:Syros Pharmaceuticals: Employment, Equity Ownership. Shearstone:Syros Pharmaceuticals: Employment, Equity Ownership.


Blood ◽  
1998 ◽  
Vol 92 (8) ◽  
pp. 2924-2933 ◽  
Author(s):  
Tohru Ikuta ◽  
Yuet Wai Kan ◽  
Paul S. Swerdlow ◽  
Douglas V. Faller ◽  
Susan P. Perrine

Abstract The mechanisms by which pharmacologic agents stimulate γ-globin gene expression in β-globin disorders has not been fully established at the molecular level. In studies described here, nucleated erythroblasts were isolated from patients with β-globin disorders before and with butyrate therapy, and globin biosynthesis, mRNA, and protein-DNA interactions were examined. Expression of γ-globin mRNA increased twofold to sixfold above baseline with butyrate therapy in 7 of 8 patients studied. A 15% to 50% increase in γ-globin protein synthetic levels above baseline γ globin ratios and a relative decrease in β-globin biosynthesis were observed in responsive patients. Extensive new in vivo footprints were detected in erythroblasts of responsive patients in four regions of the γ-globin gene promoter, designated butyrate-response elements gamma 1-4 (BRE-G1-4). Electrophoretic mobility shift assays using BRE-G1 sequences as a probe demonstrated that new binding of two erythroid-specific proteins and one ubiquitous protein, CP2, occurred with treatment in the responsive patients and did not occur in the nonresponder. The BRE-G1 sequence conferred butyrate inducibility in reporter gene assays. These in vivo protein-DNA interactions in human erythroblasts in which γ-globin gene expression is being altered strongly suggest that nuclear protein binding, including CP2, to the BRE-G1 region of the γ-globin gene promoter mediates butyrate activity on γ-globin gene expression. © 1998 by The American Society of Hematology.


Blood ◽  
1982 ◽  
Vol 59 (4) ◽  
pp. 738-746 ◽  
Author(s):  
BL Tonkonow ◽  
R Hoffman ◽  
D Burger ◽  
JT Elder ◽  
EM Mazur ◽  
...  

Abstract The human leukemia cell line, K562, produces embryonic and fetal hemoglobins and glycophorin A, proteins normally associated only with erythroid cells. Hemoglobin accumulation is enhanced by exposure of the cells to 0.05 mM hemin. We have examined K562 cells before and after exposure to hemin to determine whether expression of these erythroid proteins was shared by all cells or confined to specific subpopulations. Globin gene expression was examined by quantitation of globin mRNA sequences, using a 3H-globin cDNA molecular hybridization probe. Constitutive cells produced globin mRNA, the content of which was increased 3–4-fold by hemin. Cell-to-cell distribution of globin mRNA was determined by in situ hybridization of 3H-globin cDNA to constitutive and hemin-treated K562 cells. Virtually all cells in the culture exhibited grain counts above background, indicating globin gene expression by all cells, rather than a confined subpopulation. Virtually all hemin-treated cells had 3–5-fold higher grain counts, indicating uniformly increased globin gene expression. The glycophorin content of K562 cells was estimated by fluorescence-activated cell sorting (FACS) of cells labeled with fluorescein-labeled antiglycophorin antiserum. The vast majority of constitutive cells contained glycophorin, but exhibited to apparent increase in glycophorin accumulation after hemin exposure. Thus, glycophorin and globin genes exhibited differential responses to hemin. These differences could reflect normal differences in the patterns of specialized gene expression in stem cells. Alternatively, different aberrations of gene expression could be occurring in response to the determinants of the neoplastic properties of K562.


Blood ◽  
2004 ◽  
Vol 104 (11) ◽  
pp. 1209-1209
Author(s):  
Susanna Harju ◽  
Halyna Fedosyuk ◽  
Kenneth R. Peterson

Abstract A 213 Kb human β-globin locus yeast artificial chromosome (β-YAC) was modified by homologous recombination to delete 2.9 Kb of cross-species conserved sequence similarity encompassing the LCR 5′HS4 (Δ5′HS4 β-YAC). Three transgenic mouse lines were established; each contained two intact copies of the β-globin locus as determined by long range restriction enzyme mapping (LRRM) and Southern blot hybridization analyses. Human ε-, γ- and β-globin, and mouse α- and ζ-globin mRNAs were measured by RNAse protection in hematopoietic tissues derived from staged embryos, fetuses and adult mice. No difference in the temporal pattern of globin transgene expression was observed between Δ5′HS4 β-YAC mice and wild-type β-YAC mice. In addition, quantitative per-copy human β-like globin mRNA levels were similar between Δ5′HS4 and wild-type β-YAC transgenic lines, although γ-globin gene expression was slightly increased in the fetal liver, while β-globin gene expression was slightly decreased in Δ5′HS4 β-YAC mice. These data are in contrast to data obtained from β-YAC mice containing a deletion of the 280 bp 5′HS4 core. In these mice, γ- and β-globin gene expression was significantly decreased during fetal definitive erythropoiesis and β-globin gene expression was decreased during adult definitive erythropoiesis. However, these data are consistent with the observation that deletion of the 5′HS core elements is more deleterious than large deletions of the 5′HSs. Together, the compiled deletion data supports the hypothesis that the LCR exists as a holocomplex in which the 5′HS cores form an active site and the flanking 5′HS regions constrain the holocomplex conformation. In this model, 5′HS core mutations are dominant negative, whereas larger deletions allow the LCR to fold into alternate holocomplex structures that function normally, albeit less efficiently. To complete the study on the contribution of the individual 5′HSs to LCR function, a 0.8 Kb 5′HS1 fragment was deleted in the 213 Kb β-YAC by homologous recombination. Two ΔHS1 β-YAC transgenic lines have been established; four additional founders were recently identified. Of the two lines, one contains two intact copies of the globin locus; the other contains four deleted copies, one of which extends from the LCR through just 5′ to the β-globin gene. For both lines, ε-globin gene expression was markedly reduced, approximately 5–10 fold, during primitive erythropoiesis. Developmental expression profiles and levels of the γ- and β-globin genes (in the line that contains loci including the β-globin gene) were unaffected by deletion of 5′HS1. Breeding of the remaining four founders to obtain F1 and F2 progeny for similar structure/function studies is in progress. Decreased expression of the β-globin gene is the first phenotype ascribed to a 5′HS1 mutation, suggesting that this HS does indeed have a role in LCR function.


Sign in / Sign up

Export Citation Format

Share Document