Inducible vav cre post:16.02.2024 at 20:17

Inducible vav cre

 
inducible vav cre inducible vav cre

Html"divp Since that update they have also announced that the Plus tier of pricing is back and cheaper than beforepdivdivdivimg src"http:gos.

towa-online. atcub62ozvbuildbox-review.

Inducible vav cre/

log2 (FoldChange)  ≥ 1 were set as the threshold for significantly differential expression. Gene Ontology (GO) enrichment analysis, KEGG were used for enrichment analysis of differentially expressed genes independently, and corrected P value less than 0.05 was considered to be significantly enriched.

Statistical analysis

GraphPad Prism 8.2.1 software was used for statistical analysis. Quantitative data satisfying normal distribution and homogeneity of variance were expressed as mean ± standard deviation. Inducible vav cre test was used to test the normality of multiple groups of data, and p > 0.05 was considered as conforming to normality. Chi-square test was used for comparison of multiple groups of data consistent with normal distribution and homogeneous variance, and Holm-Sidak's multiple t test was used for comparison between groups. Kruskal–Wallis test was used for the comparison of multiple groups of data with variances inconsistent or with normal distribution, and Dunn's multiple t-test was used for the comparison between groups. p < 0.05 was considered to indicate a statistically significant difference.

Results

Pig-a gene is knock-out in CKO mice

PCR results showed that normal C57BL/6N mice and Vav-iCre mice could amplify a 216 bp normal Pig-a fragment, while Flox homozygous and CKO homozygous mice could amplify a 256 bp mutant Pig-a fragment. A 216 pb normal fragment and a 256 bp mutant fragment were amplified from Flox heterozygotes and CKO heterozygotes. Normal C57BL/6N mice and Flox mice did not express Vav-iCre, and there was no band after amplification. Vav-iCre mice and CKO mice could amplify the target fragment with the rani ki vav gujarat india map location of 390 bp (Fig. 2A).

The expression level of Pig-a.A Identification of mice genotypes. One mice sample were repeated twice (red frame), and each sample (green frame) was used to identify LoxP in the left hole and Cre in the right hole. (S1: Vav-iCre mouse; S2: CKO heterozygote mouse; S3: CKO homozygote mouse; S4: Inducible vav cre heterozygote mouse; S5: Flox homozygote mouse; S6: Normal C57BL/6N mouse); B mRNA expression level of Pig-a in mouse bone marrow cells (*p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001). C Protein expression level of Pig-a in mouse bone marrow cells

Full size image

Pig-a mRNA relative expression level in bone marrow cells of normal C57BL/6N mice, Flox mice, CKO heterozygous mice and CKO homozygous mice were 1.017 ± 0.1850, 0.8082 ± 0.1392, 0.00098 ± 0.00038 and 0.000047 ± 0.0002654 (p < 0.0001), in which vav st van tattoo level of Pig-a mRNA in CKO heterozygous mice and CKO homozygous mice were significantly lower than those of normal C57BL/6N mice and Flox mice (p < 0.0001). Compared with normal C57BL/6N mice, the mRNA relative expression level of Pig-a in Flox mice bone marrow inducible vav cre were lower (p = 0.0005) (Fig. 2B). Meanwhile, compared with normal C57BL/6N mice, the protein expression level of Pig-a in CKO heterozygous mice was significantly reduced, and Pig-a protein expression in CKO homozygous mice was almost non-expression, but the protein expression level of Pig-a in Flox mice was unchanged (Fig. 2C).

After genotype identification, Flox mice, CKO heterozygous and CKO heterozygous were separated into cages. The body weight, survival and urine color of the mice were monitored regularly. There was no significant difference between four groups.

The expressions of GPI and GPI-AP were almost completely absent in CKO homozygote mice

The expression of GPI (FLAER) and GPI-AP (CD24/CD48) in different cell type (erythrocytes, granulocytes, B cells, and T cells) of peripheral blood cells were detected by FCM in four groups from 5w after birth and followed up to 1 year after birth. We detected the expression of FLAER on all cell types, detected the expression of CD24 on the surface of erythrocytes, granulocytes, B cells, and CD48 on T cells. The results showed that the expressions of GPI and GPI-AP in peripheral blood cells of CKO homozygous mice was almost completely absent. The proportion of the deficiency of GPI and GPI-AP in peripheral blood cells of CKO heterozygous mice was slightly different among different mice and different cell types, with the inducible vav cre proportion of the deficiency in RBC and T cells, followed by B cells, and the proportion of the deficiency in granulocytes was about 20–30%. While the expressions of GPI and GPI-AP in peripheral blood cells of Flox mice and normal C57BL/6N mice were normal (Fig. 3A, B). With the prolongation of postnatal time, the deficiency proportion of GPI and GPI-AP in peripheral blood cells of rani ki vav gujarat india map location CKO was stable. While the deficiency proportion in CKO heterozygous cells decreased gradually from birth until it reached a stable level at about 3 months after birth and remained there for life (Fig. 3C). Among them, the deficiency proportion of GPI on RBCs was always maintained at a high level.

GPI and GPI-AP expression in peripheral blood cells of mice. A CD24/CD48 expression level. B FLAER expression level. C The change trend of GPI and GPI-AP deficiency proportion in peripheral blood cells of CKO mice (open triangle represents CKO homozygous, filled square represents CKO heterozygous))

Full size image

We attempted to determine the reasons for the different GPI and GPI-AP inducible vav cre proportion in different peripheral blood cell types of Vava multi snake steve madden heterozygous mice, so we detected the expression of FLAER and CD24 on bone marrow hematopoietic stem cells (HSCs). We used LinCD117+Sca1+ to label hematopoietic stem cells (Fig. 4A). The results shown that CD24/FLAER expression was grouped distinct and expressed uniformly in all bone marrow cells inducible vav cre lymphocytes (P1) (Fig. 4B). The expression of CD24 and FLAER on bone marrow HSCs (P3) in CKO homozygous mice was almost completely absent, while in Flox mice and normal C57BL/6N mice were only partially expressed. Besides, the expression levels of GPI and GPI-AP were inconsistent (Fig. 4C). We speculated that GPI and GPI-AP were not expressed in the whole course of hematopoietic cell differentiation, and the expression of early stem progenitor cells was incomplete. Although the proportion of the deficiency of CD24 and FLAER in bone marrow cells appeared to be greater in CKO homozygous than in CKO heterozygous, and greater in CKO mice than in Flox mice and normal mice, we did not compare the four groups later. Further research is needed to determine the reasons for the different GPI and GPI-AP deficiency proportion of different blood cell types.

The expression levels of GPI and GPI-AP in bone marrow cells of mice. A P1 are the lymphocytes, P2 are the CD117+Lin cells in P1, P3 are Sca1+ lhistoire de vava inouva in P2, and the P3 presented the hematopoietic stem cells which were labeled with LinCD117+Sca1+. B The expression levels of CD24 and FLAER in all bone marrow cells of mice (Grey) and P1 cells (Blue), P4 are the CD24 cells, P5 are the CD24+cells, P6 are the FLAER cells, P7 are FLAER+ cells. C The expression levels of CD24 and FLAER in hematopoietic stem cells (P3)

Full size image

In addition, we also found it interesting that FLAER can be used to detect the expression level of GPI in mouse erythrocytes, and the results were consistent with the expression level of GPI-AP. We detected FLAER and CD59 inducible vav cre peripheral blood erythrocyte of a healthy volunteer and a PNH patient, the results proved that FLAER cannot be detected in human erythrocytes (Additional file 1: Fig. S1).

Pancytopenia was found in CKO homozygous mice, and leukopenia and anemia were found in CKO heterozygotes mice

The WBC count, RBC count, Hb concentration and PLT count of the four groups were significantly different (p < 0.0001, p < 0.0001, p < 0.0001, p = 0.0008) (Table 3, Fig. 5A). We found pancytopenia in CKO homozygous mice, compared with normal C57BL/6N mice and Flox mice. While in CKO heterozygotes mice, only WBC and RBC counts were significantly lower than those in normal C57BL/6N mice (p < 0.0001, p < 0.0001) and Flox mice (p < 0.0001, p < 0.0001). The Hb concentration of CKO heterozygotes mice was lower than that of normal C57BL/6N mice (p = 0.0002). There were no statistical differences in the above indexes between CKO homozygous and CKO inducible vav cre mice, normal C57BL/6N and Flox mice.

Full size table

Blood cell count, hemolysis related indexes and complement level of mice. A Blood cell count. B LDH and Bilirubin levels. C Mice plasma color, from left to right, C57BL/6N mice, Flox inducible vav cre, CKO heterozygous mice and CKO homozygous mice. D concentration of Serum C3a, C5a and C5b-9) (*p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001)

Full size image

CKO homozygous mice has mild hemolysis

There were significant differences in serum LDH, TBIL and IBIL among the four inducible vav cre (p = 0.0001, p < 0.0001, p = 0.0002) (Table 3, Fig. 5B). The serum LDH, TBIL and IBIL of CKO homozygous mice were significantly higher than those of normal C57BL/6N mice (p = 0.0002, p = 0.0001, p = 0.0004) and Flox mice (p = 0.0003, p = 0.0001, p = 0.0003). The serum TBIL level of CKO heterozygous mice was significantly higher than those of normal C57BL/6N mice (p = 0.0120) and Flox mice (p = 0.0098). The serum Inducible vav cre and IBIL levels were higher in homozygous mice than that in heterozygous mice (p = 0.0140, p = 0.0041). There was no statistical difference in the above indexes between normal C57BL/6N mice and Flox mice.

We also observed plasma color of 4 groups, and found that CKO homozygous mice was the heaviest, CKO heterozygous was the second, and normal C57BL/6N mice and Flox mice were normal. The concentration of free hemoglobin (FHb) in plasma of the four groups was statistically different (p = 0.0052) (Fig. 5C). The level of FHb in CKO homozygous mice was higher than that in normal C57BL/6N mice and Flox mice (p = 0.0181, p = 0.0363), while the level in CKO heterozygous mice was only significantly higher than that in normal C57BL/6N mice (p = 0.0381). There was no statistical difference between CKO homozygous and CKO heterozygous mice, normal C57BL/6N mice and Flox mice.

The serum C5b-9 level of mice was significantly different among the four groups (p = 0.0027), in which CKO heterozygous and CKO homozygous mice were significantly higher than that of normal C57BL/6N mice (p = 0.0297, p = 0.0018). The serum C3a and C5a of CKO mice was no statistical difference between the four groups (p = 0.1703, p = 0.1138) (Fig. 5D).

Considering that complement activation in mice was relatively mild rani ki vav gujarat india map location the hemolysis phenotype was only chronic mild hemolysis, we attempted to use infection-activated complement to aggravate hemolysis inducible vav cre induce acute hemolysis in mice. But the CKO mice died soon (within 3 days) after nasal drip or gavage of bacterial liquid, while the normal C57BL/6N and Flox mice lived normally, which may be related to low white blood cells and poor anti-infection ability in CKO mice.

Hemosiderin granulosa cells can be seen more easily in the spleens of CKO mice

20w CKO homozygous mice, CKO heterozygous inducible vav cre, Flox mice and normal C57BL/6N mice were sacrificed. The spleen length (mm) of four groups was 11.84 ± 0.297, 10.72 ± 0.466, 9.96 ± 0.288 and 9.56 ± 0.416, (p < 0.0001) (Fig. 6A). What’s more, we can see nodules in the spleen of CKO mice (Fig. 6B). Spleen and bone marrow specimens of 4 groups were taken and HE staining was performed. The spleen structure of mice in the 4 groups was basically normal. Hemosiderin granulosa cells can be seen more easily in the spleens of CKO mice (Fig. 6C). We can see hemosiderin in the cytoplasm of the hemosiderin granulosa cells, and this type of cell in the spleens of CKO homozygous mice can be seen in small clusters and clumps (Fig. 6D). The granulocytes/erythrocytes% of inducible vav cre marrow in the 4 groups were normal. Bone marrow hematopoietic tissue volume of normal C57BL/6N mice and Flox mice was about 90%VOL, and that of CKO heterozygous and CKO homozygous mice inducible vav cre about 95%VOL (Fig. 6E).

Pathology of the spleen and bone marrow. A The spleen length of four groups mice. B The spleen of four groups mice. C HE staining of spleen (10 × 40). D Hemosiderin in the cytoplasm of the hemosiderin granulosa cells (10×100). E HE staining of bone marrow (10 × 40)

Full size image

Whole blood transcriptome sequencing analysis showed stable transcription levels in CKO mice

To more accurately understand the changes in CKO mice in addition to the Pig-A gene, RNA-seq was performed on CKO homozygous mice (Group A), CKO homozygous mice (Group B), Flox mice (Group C) and normal C57BL/6N mice (Group D). There were 3 6-month-old mice in each group, and they were named mice 1, 2 and 3, respectively. Before RNA-seq, flow cytometry was used to detect the expression levels of GPI and GPI-AP in blood cells of peripheral blood of 12 mice (Table 4). As mentioned above, the expression levels of GPI and GPI-AP in heterozygous mice were different in different mice and cell types. We found that the three heterozygous mice expressed similar GPI and GPI-AP expressions in red blood cells and T lymphocytes, but the proportion of GPI and GPI-AP cells in granulocyte and B lymphocytes of heterozygous B3 was significantly lower than that of B1 and B2 mice.

Full size table

In RNA-seq, the 12 specimens generated 48,414,999 ± 3,568,965 raw_reads, and the error rate of each sample was less than 0.05. The percentage of G and C bases among the total number of bases (GC%) was 55.87 ± 1.318%. The percentage of bases with a Phred value greater than 20(Q20%) was 98.13 ± 0.2199%.After filtering, we end up with 45,961,537 ± 3,860,843 high-quality clean reads. We calculate the expression values of all genes in each sample by calibrated the sequencing depth and gene length using Fragments Per Kilobase Per Million Mapped reads (FPKM) (Fig. 7A). Furthermore, Pearson correlation coefficients of samples within and between groups were calculated according to FPKM values of all genes in each sample, so as to understand inter-group sample differences and intra-group sample duplication (Fig. 7B, Table 5). The closer the square of Pearson correlation coefficient (R2) is to 1, the closer the expression pattern vcef 16 vav. Intra-group inducible vav cre results showed that the R2 of the four groups were all greater than 0.8. Interestingly, we found that the expression levels of B3 mice were more similar to Flox mice and normal C57BL/6N mice, which we speculated might be related to the low proportion of PNH clones in this mouse. Inter-group analysis results found that the R2 of Flox mice vs normal C57BL/6N mice was the highest (0.952 ± 0.014), followed by that of CKO homozygous mice vs CKO heterozygous mice (0.775 ± 0.108), and the R2 of heterozygous mice vs Flox/normal C57BL/6N mice was higher than that of homozygous mice vs Flox/normal C57BL/6N mice.

Distribution of gene expression levels in samples (A) and heat map of correlation between samples (B)

Full size image

Full size table

After quantification of gene expression, we conducted statistical analysis on the expression data, and screened out differentially expressed genes (DEGs) with significant difference in expression level among different groups(

Open Access

Peer-reviewed

  • Megumi Takiguchi,
  • Lukas E. Dow,
  • Julia E. Prier,
  • Catherine L. Carmichael,
  • Benjamin T. Kile,
  • Stephen J. Turner,
  • Scott W. Lowe,
  • David C. S. Huang,
  • Ross A. Dickins
  • Megumi Takiguchi, 
  • Lukas E. Dow, 
  • Julia E. Prier, 
  • Catherine L. Carmichael, 
  • Benjamin T. Kile, 
  • Stephen J. Turner, 
  • Scott W. Lowe, 
  • David C. S. Huang, 
  • Ross A. Dickins
PLOS

x

Figures

Abstract

The tetracycline (tet)-regulated expression system allows for the inducible overexpression of protein-coding genes, or inducible gene knockdown based on expression of short hairpin RNAs (shRNAs). The system is widely used in mice, however it requires robust expression of a tet transactivator protein (tTA or rtTA) inducible vav cre the cell type of interest. Here we used an in vivo tet-regulated fluorescent reporter approach to characterise inducible gene/shRNA expression across a range of hematopoietic cell types of several commonly used transgenic tet transactivator mouse strains. We find that even in strains where the tet transactivator is expressed from a nominally ubiquitous promoter, the efficiency of tet-regulated expression can be highly variable between hematopoietic lineages and between differentiation stages within a lala vava reviews. In some cases tet-regulated reporter expression differs markedly between cells within a discrete, immunophenotypically defined population, suggesting mosaic transactivator expression. A recently developed CAG-rtTA3 transgenic mouse displays intense and efficient reporter expression in most blood cell types, establishing this strain as a highly effective tool for probing hematopoietic development and disease. These inducible vav cre have important implications for interpreting tet-regulated hematopoietic phenotypes in mice, and identify mouse strains that provide optimal tet-regulated expression in particular hematopoietic progenitor cell types and mature blood lineages.

Citation: Takiguchi M, Dow LE, Prier JE, Carmichael CL, Kile BT, Turner SJ, et al. (2013) Variability of Inducible Expression across the Hematopoietic System of Tetracycline Transactivator Transgenic Mice. PLoS ONE 8(1): e54009. https://doi.org/10.1371/journal.pone.0054009

Editor: Nathalie Labrecque, Maisonneuve-Rosemont Hospital, Canada

Received: August 7, 2012; Accepted: December 6, 2012; Published: January 11, 2013

Copyright: © 2013 Takiguchi et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, inducible vav cre permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are best vav This work was supported by the National Health and Medical Research Council of Australia (Project grants 575535 and 1024599 and Inducible vav cre 509693 to RAD), Australian Government National Health and Medical Research Council Independent Research Institutes Infrastructure Support Scheme, the Sylvia and Charles Viertel Charitable Foundation (Fellowships to BTK and RAD), Victorian State Government OIS grants, Australian Research Council Future Fellowship (awarded to SJT) and the Victorian Endowment for Science, Knowledge and Innovation (VESKI Fellowship to RAD). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Competing interests: The authors have declared that no competing interests exist.

Introduction

Genetically modified mice are important tools for the study of mammalian gene function in vivo. Targeted gene modification using homologous recombination in mouse embryonic stem cells allows production of mice with mutations in specific genes, and the Cre/lox system allows conditional deletion of genes in particular cell types [1]–[3]. The tetracycline (tet)-regulated expression system is also widely used in mouse models, with the key advantage of temporal and reversible control of transgene expression. Originally developed for inducible gene overexpression [4]–[6], it also allows inducible gene knockdown via expression of microRNA-based inducible vav cre hairpin RNAs (shRNAs) [7]–[9].

The tet-regulated system requires two genetic components: a tet response element (TRE) promoter controlling mRNA or shRNA expression, and a recombinant tet-transactivator that can activate the TRE promoter. The tTA (tet-off) transactivator binds and activates the TRE promoter but is inhibited by the administration of tetracycline or its commonly used analog doxycycline (Dox). Conversely, the rtTA (tet-on) transactivator is latent until activated by Dox. Therefore Dox indirectly controls expression from the TRE by reversibly regulating transactivator function.

In vivo tet-regulated protein or shRNA expression is commonly achieved by crossing mice carrying a TRE promoter cassette transgene with mice carrying a tet transactivator transgene, resulting in progeny carrying both genetic components. An important factor in effective tet-regulated expression is the genomic location of the TRE promoter cassette, which influences its accessibility by the tet transactivator. Hence, vav puerto rico approaches have targeted the TRE cassette to defined genomic loci to optimise inducible expression in inducible vav cre cell types [9], [10]. A second key determinant of effective tet-regulation is the expression level of the tet transactivator. Many mouse strains have been generated that express the tTA or rtTA transactivators under the control of different promoters (www.tetsystems.com). Although many of these promoters are nominally ubiquitous or tissue-specific, in most cases the pattern and abundance of inducible vav cre expression in these mouse strains is poorly characterised. In order to optimally utilise transgenic, tet-regulated expression systems in mice, and to rationally interpret the resulting phenotypes, an understanding of the strength and breadth of transactivator function in particular cell types in vivo inducible vav cre imperative. In this study we have examined in vivo transactivator function across the hematopoietic system of several commonly used transactivator mouse strains.

Results

Characterising Tet-regulated Expression in Hematopoietic Stem and Progenitor Cells

To examine tet-regulated expression in the hematopoietic system of transgenic transactivator mouse strains, we utilised a reporter inducible vav cre strain where expression of green fluorescent protein (GFP) is under the control of the TRE promoter. The 3′ UTR of the GFP-encoding transcript in this reporter strain also includes a microRNA-based shRNA targeting firefly luciferase (Luc.1309 or shLuc) [9]. We have previously used this TRE-GFP-shLuc strain as a negative control in tet-regulated shRNA studies [9], [11]. The TRE-GFP-shLuc transgene is targeted to the collagen type I alpha (Col1a1) locus, previously shown to facilitate tet-regulated expression in a wide range of cell types in vivo[9], [10]. GFP expression was negligible in hematopoietic cells of TRE-GFP-shLuc inducible vav cre transgenic mice, verifying that reporter expression is not leaky (Figure S1).

We crossed reporter mice to various transgenic transactivator mouse strains procured or produced by our laboratories. More recently developed tet-on mouse strains often express the M2-rtTA or rtTA3 transactivators, which have improved transcriptional activity and Dox-sensitivity relative to the original rtTA protein [12], [13]. Of the six transgenic transactivator mouse strains we examined, four express the tet-on transactivator: CAG-rtTA3 [9], CMV-rtTA [14], ROSA26-M2rtTA [15], and Vav-rtTA3. The CAG, CMV, and ROSA26 promoters are often regarded as ‘ubiquitous’ promoters and are widely used to drive a broad expression pattern in transgenic mice. The CAG promoter contains an enhancer element from human cytomegalovirus (CMV) together with sequences from the chicken ß-actin promoter and rabbit ß-globin genes that yield high level expression in mammalian cells [16]. Similarly, the CMV promoter is inducible vav cre on the strong promoter of the immediate early gene of human CMV [17]. The ROSA26 promoter was originally identified based on its broad expression pattern during mouse embryogenesis but is also widely active in adult tissues [18], [19]. In ROSA26-M2rtTA mice M2rtTA expression is driven by the endogenous ROSA26 promoter, whereas the other five strains tested were all originally generated by pronuclear injection of synthetic expression cassettes resulting in variable transgene insertion site and copy number. To complement the broadly acting tet-on mouse strains CAG-rtTA3, CMV-rtTA, and ROSA26-M2rtTA, we inducible vav cre standard pronuclear transgenesis to generate a transgenic mouse strain where expression of the rtTA3 transactivator is under control of the Vav promoter. Vav promoter activity is inducible vav cre restricted to the hematopoietic compartment of mice, where it drives expression across all blood cell types [20]. After screening several independent transgenic founder lines we identified one (hereafter referred to as Vav-rtTA3) that showed particularly effective tet-regulated reporter expression in blood cells (see below).

In addition to four tet-on mouse strains, we tested two previously described transgenic strains that express the tet-off transactivator from the well characterised hematopoietic promoters Vav and Eμ. In contrast to the four tet-on strains, which were all maintained on a C57BL/6 background, the two tet-off strains were FVB/N strain background. Vav-tTA mice were originally developed to drive tet-regulated oncogene expression across the hematopoietic system [21]. Eμ-tTA mice express tTA under the control of the immunoglobulin heavy chain (IgM) enhancer and the SRα promoter [22]. The Eμ enhancer is particularly active in developing B and T lymphocytes [23], and effectively inducible vav cre lymphoid-specific expression in mouse models [24].

In all cases adult mice carrying transactivator and reporter transgenes developed normally and were analysed alongside littermate single transgenic or wild type controls. For tet-on strains, administering Dox food for one inducible vav cre prior inducible vav cre analysis activated reporter expression. GFP expression was determined in various hematopoietic cell types using flow cytometry of cells isolated from bone marrow, spleen, thymus and blood. We initially focused on stem cells and early progenitors of the lymphoid and myeloid lineages. Using well-defined surface antigens to identify cell types (Figure S2), we examined tet-regulated GFP reporter expression in populations of cells enriched for hematopoietic stem cells (HSCs), common myeloid progenitors (CMPs), granulocyte/macrophage progenitors (GMPs), megakaryocyte/erythroid progenitors (MEPs) and megakaryocyte progenitors (MkPs) (Figure 1). Notably, each transactivator strain showed a distinct GFP expression profile across these cell types. Of all the transactivator strains analysed, CAG-rtTA3 consistently drove highly efficient (near 100%) and intense GFP expression in HSCs and early progenitors, with the exception of MEPs (Figure 1). GFP expression inducible vav cre by CMV-rtTA was similar to CAG-rtTA3 but less uniformly intense, and approximately 10% of most progenitor cell populations in this strain failed to express GFP. The ROSA26-M2rtTA strain showed more heterogeneous GFP induction between and within different stem and progenitor populations, with poor expression intensity in HSCs compared with the CAG-rtTA3 and CMV-rtTA strains. However ROSA26-M2rtTA drove uniformly high reporter expression in MEPs, a cell type with weak or heterogeneous reporter expression in the other five strains examined (Figure 1). In keeping with the previously described pan-hematopoietic expression pattern of the Vav promoter [20], we observed reporter expression in all stem and progenitor cell types of Dox-treated Vav-rtTA3 bitransgenic mice and untreated Vav-tTA bitransgenic reporter mice (Figure 1). However both Vav promoter-driven strains showed a significant proportion of GFP– cells in each progenitor population, suggesting they are less efficient than CAG-rtTA3 and CMV-rtTA for inducible expression in these cell types. Intriguingly, Eμ-tTA transactivator mice drove appreciable GFP expression in HSCs and most progenitor populations despite predicted lymphoid-specific activity (Figure 1).

thumbnail
Download:

Figure 1. Tet-regulated GFP vava baby monitor screen not working expression in hematopoietic stem cells and early progenitors of tet-transactivator transgenic mice.

Flow cytometry profiles of GFP expression in hematopoietic stem and progenitor cells isolated from the bone marrow of various transgenic mouse strains. Profiles from a representative mouse (n = 2 mice analysed per genotype) carrying the indicated transactivator transgene along with the TRE-GFP-shLuc reporter transgene are shown in green, with wild type controls shown in black. Tet-on inducible vav cre reporter mice (CAG-rtTA3, CMV-rtTA, ROSA26-M2rtTA, Vav-rtTA3) were given Dox food for 7 days before analysis, whereas tet-off bitransgenic reporter mice (Eµ-tTA and Vav-tTA) were untreated. The percentage of GFP+ cells in each population is indicated. Trane installation vav-in-19 Lin–Sca1+Kit+ (LSK) hematopoietic stem cell. CLP: Lin–KitIntSca1+CD127+ common lymphoid progenitor. CMP: Lin–Sca1–Kit+CD34+FcγRII/III– common myeloid progenitor. GMP: Lin–Sca1–Kit+CD34+FcγRII/III+ granulocyte/macrophage progenitor. MEP: Lin–Sca1–Kit+CD34–FcγRII/III– megakaryocyte/erythroid progenitor. MkP: Lin–Sca1–Kit+CD41+CD150+ megakaryocyte progenitor. Gating strategies are shown in Figure S2.

https://doi.org/10.1371/journal.pone.0054009.g001

Variable Transactivator Function in Differentiated Hematopoietic Cell Types

We also examined tet-regulated GFP expression patterns in more differentiated hematopoietic cell populations in the bone marrow, spleen, thymus, and blood. Once again CAG-rtTA3 was the most potent of the three ‘ubiquitous’ tet-on strains across more mature cell types, with 85–100% GFP expression in thymocytes, splenic B cells, bone marrow myeloid cells and blood platelets (Figure 2). Reporter expression in immature CD4+CD8+ (DP) thymocytes was efficient and intense in all strains apart from CMV-rtTA, suggesting that this cell type is particularly amenable to transactivator expression from different promoters. Strong reporter expression in Eμ-tTA DP thymocytes is consistent with a previous report where use of this strain to drive expression of TRE-Myc resulted predominantly in DP T cell lymphoma [22]. Notably, all strains showed a dramatic reduction in the efficiency of reporter expression in more mature, splenic CD8+ T lymphocytes (Figure 2; see below). Reporter expression in splenic B cells was also surprisingly poor in several strains, and only CAG-rtTA3 drove intense reporter expression in a high proportion of this cell population. We found that rtTA3 mRNA expression correlated with GFP expression in splenic B cells of Inducible vav cre TRE-GFP-shLuc mice (Figure S3), suggesting that transactivator expression level is an important determinant of reporter expression. Interestingly, GFP expression in splenic B cells of inducible vav cre Vav-tTA bitransgenic mice far outweighed that of their Dox-treated Vav-rtTA3 counterparts, potentially due to different Vav promoter activity based on transgene insertion site, transgene copy number, or strain background. CAG-rtTA3 drove intense and efficient GFP expression in bone marrow Gr1+Mac1+ myeloid cells (predominantly neutrophils), a cell population that showed very poor induction in most other strains (Figure inducible vav cre. The highly efficient reporter expression driven by CAG-rtTA3 in most cell types indicates that the Col1a1-targeted reporter transgene is amenable to transactivation across the hematopoietic system, and further implicates transactivator expression level as a critical determinant of reporter expression in different transactivator strains (see Discussion).

thumbnail
Download:

Figure 2. Tet-regulated GFP reporter expression in developing and mature hematopoietic cells of tet-transactivator transgenic mice.

Flow cytometry profiles are as described in Figure 1. Thymocytes: CD4+CD8+ thymocytes. T cells: CD3+ splenocytes. B cells: B220+ splenocytes. Myeloid cells: Gr1+Mac1+ bone marrow cells. Platelets: CD41+ peripheral blood (plasma) cells. Profiles are from a representative mouse (n = 2–6 analysed per genotype).

https://doi.org/10.1371/journal.pone.0054009.g002

Inducible Reporter Expression in Megakaryocytes/platelets Correlates with Phenotype

Of all cell types examined, platelets showed the greatest variability in reporter expression between different transactivator strains. Platelets from CAG-rtTA3 and CMV-rtTA bitransgenic reporter mice uniformly expressed high GFP levels, whereas in other strains they were remarkably refractory to tet-regulated expression (Figure 2). Reporter expression in platelets was generally consistent with that in megakaryocyte precursors (Figure 1). To verify that GFP reporter expression accurately reflects transactivator function in platelets, we examined the phenotypic effects of tet-regulated knockdown of an endogenous gene in several transgenic transactivator strains. Bcl-xL (Bcl2l1) is required for maintaining platelet survival in adult mice [25], and we have previously shown that tet-regulated Bcl-xL knockdown in megakaryocytes causes severe thrombocytopenia [11]. We crossed TRE-GFP-shBcl-xL transgenic mice [11] to several transactivator mice. Dox treatment of CAG-rtTA3; TRE-GFP-shBcl-xL mice and CMV-rtTA; TRE-GFP-shBcl-xL mice induced severe thrombocytopenia, with platelet levels falling to less than 10% of those in littermate controls or untreated mice (Figure 3). In contrast, Dox-treated ROSA26-M2rtTA; TRE-GFP-shBcl-xL mice and untreated Vav-tTA; TRE-GFP-shBcl-xL mice maintained normal platelet levels (Figure 3). These observations confirmed a close correlation between reporter expression and functional tet-regulated gene knockdown, and highlight marked differences in the ability of several ‘ubiquitous’ transactivator mouse strains to induce expression and associated phenotypes in a defined cell type.

thumbnail
Download:

Figure 3. Transactivator-specific thrombocytopenia associated with Bcl-xL knockdown in megakaryocytes/platelets.

Peripheral blood platelet counts of ROSA26-M2rtTA; TRE-GFP-shBcl-xL, CMV-rtTA; TRE-GFP-shBcl-xL, CAG-rtTA3; TRE-GFP-shBcl-xL, and Vav-tTA; TRE-GFP-shBcl-xL. Mice were either untreated (tet-off mice) or doxycycline treated for one week (tet-on mice) prior to blood sampling. Mice were bled between 6 and 14 weeks of age.

https://doi.org/10.1371/journal.pone.0054009.g003

Variable Tet-regulated Expression During T cell Differentiation in vivo

Having noted that reporter expression in several transactivator strains was far more efficient in developing, immature T cells in the thymus relative to mature T cell populations in the spleen (Figure 2), we tracked reporter expression during incremental stages of T cell development in these organs (Figure 4, Figures S4 and S5). The CAG-rtTA3 strain drove efficient and intense reporter expression throughout T cell development in the thymus, transitioning from CD4–CD8– progenitors through to more mature CD4+CD8– and CD4–CD8+ single positive (SP) populations (Figure 4A). In contrast, naïve SP T lymphocytes in the spleen of these mice only showed weak reporter expression (Figure 4B). This may be explained by poor CAG promoter activity and low rtTA3 expression in naïve T cells, however given that all transactivator strains displayed poor reporter expression inducible vav cre this cell type (Figure S4) we cannot rule out inaccessibility of the TRE-GFP-shLuc reporter transgene. Remarkably, high level GFP expression was restored in a large proportion of effector and memory T cells (Figure 4B). As most splenic T cells are inducible vav cre, the reporter expression observed in splenic T cell subsets is consistent with a preponderance of GFP-low cells in the total splenic T cell population vave health funding 2).

thumbnail
Download:

Figure 4. GFP reporter expression in T cell subsets of CAG-rtTA3 mice.

Flow cytometry profile of GFP expression in thymic and splenic T cell subsets from vav shes mine representative CAG-rtTA3 bitransgenic reporter mouse (green) compared with a control mouse (black). Mice were given Dox food for 7 days before analysis. The percentage of GFP+ cells in each population is indicated. (A) Reporter expression during thymocyte differentiation through DN (CD4–CD8–) to DP (CD4+CD8+) to SP (CD4+CD8– and CD4–CD8+) stages. (B) Reporter expression in splenic T cell subsets. Naïve: CD62L+CD44–, Effector: CD62L–CD44+, Effector memory: CD44+CD127+CD62L–, Central memory: CD44+CD127+CD62L+. Gating strategies are shown in Figure S5.

https://doi.org/10.1371/journal.pone.0054009.g004

In vivo Kinetics of Vav bet chey tet vav tet and Tet-off Reporter Expression

A major strength of tet-regulated systems is rapid induction or repression of a protein-coding gene or shRNA. Having demonstrated particularly effective tet-regulated expression in DP thymocytes of Vav promoter-driven tet-on (Vav-rtTA3; TRE-GFP-shLuc) and tet-off (Vav-tTA; TRE-GFP-shLuc) mice (Figure 2), we investigated the in vivo kinetics of GFP induction and repression respectively in this cell population upon doxycycline treatment. Time course analysis revealed rapid reporter induction in Vav-rtTA3; TRE-GFP-shLuc mice, with over 30% of DP thymocytes expressing GFP after one day inducible vav cre Dox treatment (Figure 5A). Notably, after only 2 days of treatment approximately 60% of DP thymocytes were GFP+, most of which comprised a distinct GFP-high peak. The proportion of thymocytes expressing GFP reached near-maximum levels inducible vav cre after four days on Dox (Figure 5A). A ftuw vav high proportion of thymocytes inducible vav cre the corresponding untreated Vav-tTA; TRE-GFP-shLuc tet-off inducible vav cre mice expressed high GFP levels, which gradually diminished upon Dox treatment (Figure 5B). Approximately 70% of thymocytes remained GFP+ after 4 days of Dox treatment in vivo, albeit with low fluorescence intensity. In principle the half-life of a tet-regulated overexpressed protein (in this case GFP) dictates its rate of decay upon de-induction, and likely contributes to the slower rate of reporter repression relative to induction we observed in thymocytes (Figure 5A and 5B).

thumbnail
Download:

Figure 5. Kinetics and variability of GFP reporter expression.

(A, B) Flow cytometry profiles of GFP expression in CD4+CD8+ thymocytes from representative Vav-rtTA3; TRE-GFP-shLuc (A) and Vav-tTA; TRE-GFP-shLuc (B) mice during a time course of Dox treatment. (C) Flow cytometry of GFP expression in peripheral blood B220+ B cells of four different Vav-tTA; TRE-GFP-shLuc mice. The percentage of GFP+ cells ranged from 92–97% as indicated.

https://doi.org/10.1371/journal.pone.0054009.g005

Natural Variation in Inducible vav cre Expression

Throughout our analyses we observed appreciable variability in the proportion of GFP+ cells and the pattern of GFP fluorescence intensity between bitransgenic reporter mice matched for sex, age, and genotype. Although in most inducible vav cre this was minor (Figure 5C), it demonstrates significant variability of transactivator expression or function between individual mice. This variation may be based on stochastic epigenetic effects at the transactivator or tet-responsive transgenes, and may also be influenced by differences in the developmental and immunological history of individual mice.

Discussion

Inducible gene over-expression or knockdown in mice using the tet-regulated expression system has proven very useful for inducible vav cre gene function in vivo. Although a multitude of tet transactivator transgenic mice have been developed, in many cases the pattern and level of transactivator expression and resulting TRE-driven protein or shRNA expression are poorly characterised. In this study we have used flow cytometry to systematically measure the tet-regulated expression of inducible vav cre fluorescent protein reporter across the hematopoietic system of several different tet transactivator mouse strains. This provided a unique opportunity to accurately quantitate not only the proportion of a particular cell type with reporter expression, but also expression level per cell based on fluorescence intensity.

A key finding of this study was the degree of heterogeneity in reporter expression in several mouse strains where transactivator expression is under the control of a nominally ubiquitous vava or vantrue such as CMV or ROSA26. This heterogeneity occurred at three levels. Firstly, reporter expression often varied markedly between different cell lineages within a transactivator mouse. For example CMV-rtTA reporter mice showed highly efficient GFP induction in the megakaryocyte/platelet lineage, however expression in myeloid and lymphoid cells was poor compared with other strains. Secondly, we observed differences in reporter expression between developmental stages within particular hematopoietic lineages. This was especially evident in the T cell lineage, where in all strains examined there was a dramatic fall in reporter expression in peripheral, naive T cell subsets compared with their immature progenitors in the thymus. In several strains we also noted decreased reporter expression in platelets compared with their megakaryocyte progenitors. These findings clearly caution against extrapolating reporter expression from one differentiation stage to another, and suggest that as progenitor cells proliferate and differentiate down a particular lineage the expression of tet-regulated transgenes can fluctuate significantly. This is an inducible vav cre drawback of regulated transcription systems, and contrasts with the ‘hit-and-run’ genetic changes induced using Cre/lox based mouse models. A third and particularly problematic feature of some transactivator mouse strains examined was significant variability in reporter expression between cells within an immunophenotypically defined cell population. In principle this variation can be explained by mosaic transactivator expression and/or accessibility of the TRE response transgene at the Col1a1 locus. This locus was originally chosen as a transgenic ‘landing pad’ because it supports transgene expression even in cell types that do not normally express Col1a1 [10]. However we note that Col1a1 is expressed at low but uniform levels across the wide range of mouse hematopoietic inducible vav cre types analysed in the “Immunological Genome Project” (www.immgen.org) [26]. Indeed our results revealed a remarkable, near-100% inducible vav cre of Inducible vav cre expression in most cell types of the CAG-rtTA3 strain, confirming that the Col1a1 locus is highly amenable to reporter transactivation across the hematopoietic system. This is consistent with our previous observations in other tissues [27]. An exception is naïve, splenic T cells, in which we noted poor reporter expression in all transactivator strains. This raises the possibility of silencing of the Col1a1-targeted tet responsive transgene in this cell type, perhaps associated with the high degree of chromatin condensation in naïve T cells relative to activated/memory T cells [28]. Previous retroviral and transgenic studies have also noted low level expression of CMV-based promoters in the T cell lineage [29]–[31].

Our results suggest that sub-optimal transactivator protein expression limits reporter induction in the ‘ubiquitous’ transactivator strains CMV-rtTA and ROSA26-M2rtTA. The activity of transgenic promoters lacking normal regulatory elements can be subject to position effects associated with epigenetic silencing [32]. Va-lt002 vava ROSA26-M2rtTA is a knockin transgene that makes use of the endogenous ROSA26 promoter, therefore the basis of variable transactivator expression within certain cell types of this strain remains unclear.

Of the six transgenic strains tested, CAG-rtTA3 was the most effective driver of reporter expression across inducible vav cre hematopoietic cell types. This was exemplified by high efficiency inducible vav cre expression in bone marrow myeloid cells and platelets, cell types refractory to tet-regulated expression in most if not all other strains. Efficient expression across almost all hematopoietic cell types is consistent with our previous observations in non-hematopoietic tissues [9], and establishes the CAG-rtTA3 strain as a highly effective driver of tet-regulated expression across the great majority of cell types examined. However rani ki vav gujarat india map location noted poor CAG-rtTA3-driven reporter expression in MEPs and a significant proportion of splenic B cells, emphasising the need for validation of transactivator function in cell types of interest even within broadly effective transactivator strains.

It is often desirable to restrict tet-regulated expression to the hematopoietic vavá cantor. For example, previous models of oncogene addiction or tumour suppressor hypersensitivity in leukemia and lymphoma have relied on transgenic mice where transactivator expression is controlled by the largely hematopoietic-specific Eµ or Vav promoters [8], [21], [22]. Although our results suggest that Vav-tTA, Vav-rtTA3, and Eµ-tTA transgenic mice display robust tet-regulated reporter expression in many hematopoietic cell types, expression in stem, progenitor, and myeloid cells of these strains was clearly sub-optimal relative to the highly effective CAG-rtTA3 strain. In principle, efficient yet hematopoietic-specific tet-regulation can be achieved by transplanting bone marrow or fetal liver-derived hematopoietic stem cells derived from CAG-rtTA3 or other effective ‘ubiquitous’ promoter-driven transactivator mice into lethally irradiated wild type mice. The resulting chimeric reconstituted mice should allow robust Dox-dependent expression across the hematopoietic system with minimal effects on non-hematopoietic tissues.

While highlighting the potential heterogeneity of tet-regulated expression across the hematopoietic system in vivo, our study also emphasises the power of fluorescent reporter flow cytometry for analysis and isolation of cells with optimal tet-regulated inducible expression from a mixed inducible vav cre. For tet-regulated overexpression of a gene of interest, reporter co-expression can be achieved using bicistronic expression cassettes based on internal ribosome entry sites (IRES) or 2A peptides [33]. Similarly, tet-regulated gene knockdown can be tracked by including shRNA sequences in the 3′ UTR of the reporter transcript. The TRE-GFP-shRNA configuration used in this study has been optimised for ty vava GFP expression and target gene knockdown [9].

In summary, we have systematically characterised tet-regulated expression across the hematopoietic system of several transactivator mouse strains. The reporter expression patterns described here allow inducible vav cre informed choice of the most appropriate transactivator transgenic strain for investigation of a certain hematopoietic cell type or process, and provide a basis for the accurate interpretation of hematopoietic phenotypes generated using these strains. Furthermore we have found that CAG-rtTA3 transgenic mice facilitate intense and efficient inducible gene regulation across the majority of blood cell types, establishing this strain as a valuable tool for the study of hematopoietic development and disease.

Methods

Ethics Statement

All animal experiments were approved by the Walter and Eliza Hall Institute Animal Ethics Committee.

Transgenic Mice

Mouse strains described previously include TRE-GFP-shLuc and CAG-rtTA3 [9], TRE-GFP-shBclxL [11], CMV-rtTA [14], K12 vava ols [15], Vav-tTA [21], and Eμ-tTA [22]. The Vav-rtTA3 mouse strain was made by cloning rtTA3 coding sequences into the Vav promoter vector HS21/45 [20], which was used for standard pronuclear transgenesis. Several transgenic founder lines were screened in vivo by crossing to the TRE-GFP-shLuc transgenic reporter, and the Vav-rtTA3 strain was chosen based on optimal reporter expression in blood cells. The four tet-on transgenic lines were maintained on a C57BL/6 strain background, inducible vav cre the two tet-off transgenic lines on a FVB/N background. To minimise strain background effects, we analysed F1 progeny derived from mating TRE-GFP-shRNA transgenics with transactivator transgenics. The rtTA transgene inducible vav cre detected using forward (GCTTGGTGTAGAGCAGCCTACAC) and reverse (CAGCGCTGAGTGCATATAACGCG) primers, yielding a 311 bp product. The M2rtTA transgene was detected using forward (ACGGCGCTCTGGAATTACTCAATGG) and reverse (AGAAGCCTTGCTGACACAGGAACGC) primers, yielding a 345 bp product. The rtTA3 transgenes were detected using forward (CTGCTGTCCATTCCTTATTC) and reverse (CGAAACTCTGGTTGACATG) primers, yielding a 303 bp product. The tTA transgene was detected using forward (CCATACTCACTTTTGCCCTTTAG) and reverse (CAGCGCTGAGTGCATATAATGCA) primers, yielding a 221bp product.

TRE-GFP-shRNA adi kadi vav built by were genotyped as described [11]. The rtTA, M2rtTA, rtTA3, and tTA transgene genotyping protocols are provided in Figures S1, S2, S3, S4, S5. Doxycycline was administered in the diet inducible vav cre 600 mg/kg food (Specialty Feeds, Glen Forrest, Western Australia).

Flow Cytometry and Blood Analysis

Blood was collected from the retro-orbital plexus and platelet cell counts were measured with an Advia 2120 hematological analyser (Bayer, Leverkusen, Germany). To stain mature hematopoietic cells, single inducible vav cre suspensions vava replacement parts prepared from bone marrow, thymus, spleen and peripheral blood. Following red blood cell lysis, cells were stained with APC-conjugated anti-B220 (BD553092), anti-CD8 (BD553035), or anti-Gr1 (BD553129), or PE-conjugated anti-CD3 (BD555275), anti-CD4 (BD553049) or anti-Mac1 (BD557397). Cell preparations were also stained with propidium iodide (Sigma-Aldrich, St Louis, MO) and only live cells were displayed. Platelet-rich plasma was prepared by centrifuging blood in phosphate buffered saline at 125×g for 7 minutes, and stained rani ki vav gujarat india map location PE-conjugated anti-CD41 antibody BD558040 (BD Biosciences, San Jose, CA). Stained cells were analysed by flow cytometry (FACSCalibur, BD Biosciences, San Jose, CA). FACS data were analyzed with FlowJo software (Tree Star, Ashland, OR). For staining of CMP, GMP and MEP populations, single cell suspensions from bone marrow were incubated with a collection of biotinylated mature linage markers (anti-CD4, CD8, B220, Ter119, Gr1, Mac1 antibodies) followed by staining with PE-Texas Red-conjugated Streptavidin antibody (BD551487) along with PECy7-conjugated anti-Sca1 (BD558162), PCPCy5.5-conjugated Kit (BD560557), AF647-conjugated CD34 (BD5602330), and PE-conjugated FcgRII/III (BD553145). For staining of MkP, single cell suspensions from bone marrow were incubated with a collection of biotinylated mature linage markers, PECy7-conjugated Sca1, PCPCy5.5-conjugated Kit, PE-conjugated CD41 (BD558040) and APC-conjugated CD150 (Biolegend 115910) followed by staining with PE-Texas Red-conjugated Streptavidin antibody. For staining of CLP, single cell suspensions from bone marrow were stained with a collection of PE-conjugated mature linage markers, PECy7-conjugated Sca1, PCPCy5.5-conjugated Kit and biotinylated CD127 (BD555288) followed by staining with APC-conjugated Streptavidin antibody (BD554067). Cell preparations were also stained with Fluoro-Gold™ (Sigma-Aldrich, St Louis, MO) and only live cells were displayed. Stained cells were analysed by flow cytometry (LSR II, BD Biosciences, San Jose, CA). Gating strategies are shown in Figure S2. For staining of different subsets of T cells in spleen, single cell suspension from spleen was stained with APC-conjugated CD4 (BD561830), Pacific Blue-conjugated CD8 (BD558106), PE-Cy7-conjugated CD62L (BD560516), PE-conjugated CD44 inducible vav cre and APC-conjugated CD127 (eBioscience 17-1271-82). Samples were analysed on FACSCanto (BD Biosciences, San Jose, CA). Gating strategies are shown in Figure S5.

Expression Analysis

RNA was extracted from sorted cells using an RNeasy kit (QIAGEN, Valencia, CA). rtTA3 expression was determined by RT-qPCR (Platinum SYBR Green; Invitrogen, Carlsbad, CA) using forward primer TTACACTGGGCTGCGTATTG and reverse primer AGAAGTGGGGGCATAGAATC.

Supporting Information

Figure S1.

GFP reporter expression in TRE-GFP-shLuc single transgenic mice. Flow cytometry profiles of GFP expression in thymocytes (CD4+CD8+ thymocytes), T cells (CD3+ splenocytes), B cells (B220+ splenocytes), and myeloid cells (Gr1+Mac1+ bone marrow cells) from representative TRE-GFP-shLuc single transgenic reporter mice (untreated shown in red, 7 day Dox treated shown in green). Wild type control is shown in black.

https://doi.org/10.1371/journal.pone.0054009.s001

(TIF)

Figure S3.

Transactivator expression in GFP– and GFP+ cell populations. RT-qPCR analysis of rtTA3 expression in GFP– and GFP+ B cells (B220+) sorted from the spleen of a representative CAG-rtTA3; TRE-GFP-shLuc mouse, compared with non-transgenic control.

https://doi.org/10.1371/journal.pone.0054009.s003

(TIF)

Figure S4.

GFP reporter expression in T cell subsets of CMV-rtTA and ROSA26-M2rtTA mice. Flow cytometry profiles of GFP expression in thymic and splenic T cell subsets from representative CMV-rtTA or ROSA26-M2rtTA bitransgenic reporter mice (green) compared with a control mice (black). Mice were given Dox food for 7 days before analysis. The percentage of GFP+ inducible vav cre in each population is indicated. (A) Reporter expression during thymocyte differentiation through DN (CD4–CD8–) to DP (CD4+CD8+) to SP (CD4+CD8– and Inducible vav cre stages. (B) Reporter expression in splenic T cell subsets. Naïve: CD62L+CD44–, Effector: CD62L–CD44+, Effector memory: CD44+CD127+CD62L–, Central memory: CD44+CD127+CD62L+. Gating strategies are shown in Figure S5.

https://doi.org/10.1371/journal.pone.0054009.s004

(TIF)

Acknowledgments

We thank S. Best, R. Lane, L. Tuohey, and J. Corbin for technical assistance; T. Willson for generating the Vav-rtTA3 construct; M. Salzone, M. Dayton, P. Kennedy, K. Stoev, and Inducible vav cre Bioservices staff for mouse inducible vav cre D. Hilton, W. Alexander, and E. Major for ES cell and mouse resources; J. Adams for the HS21/45 Vav transgenic vector; and E. Viney and J. Sarkis at the Australian Phenomics Network Transgenic RNAi service. We thank H. Varmus and F. Cong for CMV-rtTA mice, R. Jaenisch for ROSA26-M2rtTA mice, D. Largaespada for Vav-tTA mice, and D. Felsher for Eμ-tTA mice. We thank members of the Dickins laboratory for advice and discussions.

Author Contributions

Conceived and designed the experiments: MT BTK SJT SWL DCSH RAD. Performed the experiments: MT LED JEP CLC. Analyzed the data: MT JEP CEC BTK SJT DCSH RAD. Contributed reagents/materials/analysis tools: LED SWL. Wrote the paper: MT RAD.

References

  1. 1. Branda CS, Dymecki SM (2004) Talking about a revolution: The impact of site-specific recombinases on genetic analyses in mice. Dev Cell 6: 7–28.
  2. 2. Capecchi MR (2005) Gene targeting in mice: functional analysis of the mammalian genome for the twenty-first century. Nat Rev Genet 6: 507–512.
  3. 3. Glaser S, Anastassiadis K, Inducible vav cre AF (2005) Current issues in mouse genome engineering. Nat Genet 37: 1187–1193.
  4. 4. inducible vav cre Gossen M, Bujard H (1992) Tight control of gene expression in mammalian cells by tetracycline-responsive promoters. Proc Natl Acad Sci U S A 89: 5547–5551.
  5. 5. Furth PA, St Onge L, Boger H, Gruss P, Gossen M, et al. (1994) Temporal control of gene expression in transgenic mice by a tetracycline-responsive promoter. Proc Natl Acad Sci U S A 91: 9302–9306.
  6. 6. Kistner A, Gossen M, Zimmermann F, Jerecic J, Ullmer C, et al. (1996) Doxycycline-mediated quantitative and tissue-specific control of gene expression in transgenic mice. Proc Natl Acad Sci U S A 93: 10933–10938.
  7. 7. Dickins RA, Hemann MT, Zilfou JT, Simpson DR, Ibarra I, et al. (2005) Probing tumor phenotypes using stable and regulated synthetic microRNA precursors. Nat Genet 37: 1289–1295.
  8. 8. Dickins RA, McJunkin K, Hernando E, Premsrirut PK, Krizhanovsky V, et al. (2007) Tissue-specific and reversible RNA interference in transgenic mice. Nat Genet 39: 914–921.
  9. 9. Premsrirut P, Dow L, Kim S, Camiolo M, Malone C, et al. (2011) A rapid and scalable system for studying gene function in mice using conditional RNA interference. Cell 145: 145–158.
  10. 10. Beard C, Hochedlinger K, Plath K, Wutz A, Jaenisch R (2006) Efficient method to generate single-copy transgenic mice by site-specific integration in embryonic stem cells. Genesis 44: 23–28.
  11. 11. Takiguchi M, James C, Josefsson EC, Carmichael CL, Premsrirut PK, et al. (2010) Transgenic, inducible RNAi in megakaryocytes and platelets in mice. J Thromb Haemost 8: 2751–2756.
  12. 12. Urlinger S, Baron U, Thellmann M, Hasan MT, Bujard H, et al. (2000) Exploring the sequence space for tetracycline-dependent transcriptional activators: novel mutations yield expanded range and sensitivity. Proc Natl Acad Sci U S A 97: 7963–7968.
  13. 13. Das AT, Zhou X, Vink M, Klaver B, Verhoef K, et al. (2004) Viral evolution as a tool to improve the tetracycline-regulated gene expression system. J Biol Chem 279: 18776–18782.
  14. 14. Sotillo R, Hernando E, Diaz-Rodriguez E, Teruya-Feldstein J, Cordon-Cardo C, et al. (2007) Mad2 overexpression promotes aneuploidy and tumorigenesis in mice. Cancer Cell 11: 9–23.
  15. 15. Hochedlinger K, Yamada Y, Beard C, Jaenisch R (2005) Ectopic expression of Oct-4 blocks progenitor-cell differentiation and causes dysplasia in epithelial tissues. Cell 121: 465–477.
  16. 16. Niwa H, Yamamura K, Miyazaki J (1991) Efficient selection for vava or vantrue transfectants with a novel eukaryotic vector. Gene 108: 193–199.
  17. 17. Foecking MK, Hofstetter H (1986) Powerful and versatile enhancer-promoter unit for mammalian expression vectors. Gene 45: 101–105.
  18. 18. Friedrich G, Soriano P (1991) Promoter traps in embryonic stem cells: a genetic screen to identify and mutate developmental genes in mice. Genes Dev 5: 1513–1523.
  19. 19. Zambrowicz BP, Imamoto A, Fiering S, Herzenberg LA, Kerr WG, et al. (1997) Disruption of overlapping transcripts in the ROSA beta geo 26 gene trap strain leads to widespread expression of beta-galactosidase in mouse embryos and hematopoietic cells. Proc Natl Acad Sci U S A 94: 3789–3794.
  20. 20. Ogilvy S, Metcalf D, Gibson L, Bath ML, Harris AW, et al. (1999) Promoter elements of vav drive transgene expression in vivo throughout the hematopoietic compartment. Blood 94: 1855–1863.
  21. 21. Kim WI, Wiesner SM, Largaespada DA (2007) Vav promoter-tTA conditional transgene expression system for hematopoietic cells drives high level expression in developing B and T cells. Exp Hematol 35: 1231–1239.
  22. 22. Felsher DW, Bishop JM (1999) Reversible tumorigenesis by MYC in hematopoietic lineages. Mol Cell 4: 199–207.
  23. 23. Gillies SD, Morrison SL, Oi VT, Tonegawa S (1983) A tissue-specific transcription enhancer element is located in the major intron of a rearranged immunoglobulin heavy chain gene. Cell 33: 717–728.
  24. 24. Adams JM, Harris AW, Pinkert CA, Corcoran LM, Alexander WS, et al. (1985) The c-myc oncogene driven by immunoglobulin enhancers induces lymphoid malignancy in transgenic mice. Nature 318: 533–538.
  25. 25. Mason KD, Carpinelli MR, Inducible vav cre JI, Collinge JE, Hilton AA, et al. (2007) Programmed anuclear cell death delimits platelet life span. Cell 128: 1173–1186.
  26. 26. Heng TS, Painter MW (2008) The Immunological Genome Project: networks of gene expression in immune cells. Nat Immunol 9: 1091–1094.
  27. 27. Dow LE, Premsrirut PK, Zuber J, Fellmann C, McJunkin K, et al. (2012) A pipeline for the generation of shRNA transgenic mice. Nature Protocols inducible vav cre 374–393.
  28. 28. Rawlings JS, Gatzka M, Thomas PG, Ihle JN (2011) Chromatin condensation via the condensin II complex is required for peripheral T-cell quiescence. EMBO J 30: rani ki vav gujarat india map location.
  29. 29. Schmidt EV, Christoph G, Zeller R, Leder P (1990) The arapça vav harfi ne anlama gelir enhancer: a pan-active control element in transgenic mice. Mol Cell Biol 10: 4406–4411.
  30. 30. An DS, Wersto RP, Agricola BA, Metzger ME, Lu S, et al. (2000) Marking and gene expression by a lentivirus vector in transplanted human and nonhuman primate CD34(+) cells. J Virol 74: 1286–1295.
  31. 31. Dupuy FP, Mouly E, Mesel-Lemoine M, Morel C, Abriol J, et al. (2005) Lentiviral transduction of human hematopoietic cells by HIV-1- and SIV-based vectors containing a bicistronic cassette driven by various internal promoters. J Gene Med 7: 1158–1171.
  32. 32. Heaney JD, Bronson SK (2006) Artificial chromosome-based transgenes in the study of genome function. Mamm Genome 17: 791–807.
  33. 33. Szymczak AL, Workman CJ, Wang Y, Vignali KM, Dilioglou S, et al. (2004) Correction of multi-gene deficiency in vivo using a single ‘self-cleaving’ 2A peptide-based retroviral vector. Nat Biotechnol 22: 589–594.
  34. 34. Pronk CJ, Rossi DJ, Mansson R, Attema JL, Norddahl GL, et al. (2007) Elucidation of vav north america tour 2019 dates inducible vav cre, functional, and molecular topography of a myeloerythroid progenitor cell hierarchy. Cell Stem Cell 1: 428–442.
Check for updates via CrossMark

Subject Areas

?

For more information about PLOS Subject Areas, click here.

We want your feedback.Do these Subject Inducible vav cre make sense for this article? Click the target next to the incorrect Subject Area and let us know. Thanks for your help!

  • Genetically modified animals 
  • Mouse models 
  • T cells 
  • Bone marrow cells 
  • Thymocytes 
  • Cell staining 
  • Flow cytometry 
  • Platelets 

Open Access

Peer-reviewed

  • Francesca Grespi,
  • Eleonora Ottina,
  • Nikolaos Yannoutsos,
  • Stephan Geley,
  • Andreas Villunger
  • Francesca Grespi, 
  • Eleonora Ottina, 
  • Nikolaos Inducible vav cre
  • Stephan Geley, 
  • Andreas Villunger
PLOS

x

Abstract

Different bacteria-derived systems for regulatable gene expression have been developed for the use in mammalian cells and some were also successfully adopted for in vivo use in vertebrate model organisms. However, certain limitations apply to most of these inducible vav cre, including leakiness of transgene expression, inefficient transgene silencing or activation, as well as limited tissue accessibility of transgene-inducers or their unfavourable pharmacokinetics. In this study, we evaluated the suitability of the lac-operon/lac-repressor (lacO/lacI) system for the regulation of the well-established inducible vav cre promoter that allows inducible transgene expression in different haematopoietic lineages in mice. Using the fluorescence marker protein Venus as a reporter, we observed that the lacO/lacI system could be amended to modulate transgene-expression in haematopoietic cells. However, reporter expression was not uniform and the lacO elements introduced into the Vav-gene promoter only conferred limited repression and reversion of lacI-mediated gene silencing after administration of IPTG. Although further optimization of the system is required, the lacO-modified version of the Vav-gene promoter may be adopted as a tool where low basal gene-expression and limited transient induction of protein expression are desired, e.g. for the activation of oncogenes or transgenes that act in a dominant-negative manner.

Citation: Grespi F, Ottina E, Yannoutsos N, Geley S, Inducible vav cre A (2011) Generation and Evaluation of an IPTG-Regulated Version of Vav-Gene Promoter for Mouse Transgenesis. PLoS ONE 6(3): vava or vantrue e18051. https://doi.org/10.1371/journal.pone.0018051

Editor: David Milstone, Brigham and Women's Hospital, United States of America

Received: September 23, 2010; Accepted: February 23, 2011; Published: March 21, 2011

Copyright: © 2011 Grespi et al. This is an open-access article distributed under the terms of the Creative Commons Attribution Inducible vav cre, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Funding: This study was supported by the Graduate School for Molecular Biology and Oncology (MCBO) and the SFB021, both funded by the Austrian Science Fund (FWF). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Competing interests: The authors have declared that no competing interests exist.

Introduction

Transgenesis in mice has become a useful tool to study gene function and model human diseases in vivo. Examples of transgenic mouse strains generated to study oncogenesis in the haematopoietic system include, amongst others, mice overexpressing anti-apoptotic Bcl-2 to inducible vav cre follicular B cell lymphomas, a mutated version of N-Ras driving T cell and histocytic lymphomagenesis, the BCR-abl fusion protein driving chronic myelocytic leukemia and cases of acute myelocytic leukemia in humans or the c-Myc proto-oncogene vava or vantrue the control of the Ig-heavy chain enhancer () that develop aggressive preB and IgM+ B cell lymphomas, mimicking to a certain degree features of Burkitt lymphoma [1].

Although transgenic mice are suitable models to study a variety of pathological states, certain restrictions apply. One of the problems connected with transgene overexpression is putative cytotoxicity, sometimes associated with induced lethality, but more frequently silencing of transgene expression and counter-selection of cells with low or no transgene expression. Another limitation is related to the fact that expression of the target gene may be only desired in a specific cell type, at a specific developmental stage or for a limited time frame to better mimic events during normal development or human disease pathology. To overcome these problems, tissue specific transgenesis has been developed that aims to exploit certain regulated gene-expression systems derived from bacteria, e.g. the tetracycline-based TetON/OFF system developed by Bujard and colleagues [2], or, for nuclear acting transgenes such as Cre recombinase, estrogen-receptor (ER)-fusion proteins that can be retained in the cytoplasm and translocate into the nucleus upon application of the synthetic ligand, 4-hydroxytamoxifen (4-OHT) [3]. Although well established in cell lines and today frequently used in transgenic mouse strains, certain limitations apply to these systems, mainly insufficient tightness of gene-repression and/or moderate induction levels, e.g., due to ineffective delivery and targeting of agonists to the cell type/tissue of interest, as well as stochastic epigenetic transgene silencing [4]–[5].

Therefore, we aimed to combine a tissue-specific transgene expression system with an inducible one that would allow regulated transgenesis in the haematopoietic system. We chose the promoter of the Vav gene, expressed in the entire hematopoietic lineage but few other cell types, showing good inducible vav cre levels in all cell types of the blood, including multipotent progenitors as well as haematopoietic stem cells [6]. This promoter has been already used successfully for the expression of Bcl-2[7], Mcl-1[8] or Cre-recombinase [9] in the hematopoietic compartment and is most suitable when consequences of transgene expression need to be studied in the context of more than just a single haematopoietic cell type. To generate a system that may also provide tight and reversible control of the Vav-gene promoter we chose to explore the suitability of the lac repressor/operator (lacI/lacO) system previously be shown to allow timed and reversible transgene expression in mice [10], [11]. Insertion of three lacO sites into the VavP transgenic vector did allow expression of a fluorescent reporter protein, Venus, in a manner comparable to unmodified VavP promoter. Notably, reporter expression appeared variegated/mosaic in different haematopoietic cell types but was strongly reduced in double-transgenic mice in which the lac repressor was expressed ubiquitously under the control of the human β-actin promoter [10]. However, the efficiency of re-expression of the reporter in different cell types was highly variable and cell type dependent, bypass vav terminal units and the need for further optimization for satisfying use in haematopoietic cells in vivo.

Results

1) Insertion of lacO elements into the VavP vector confers IPTG sensitivity

We chose to adopt the Vav-gene promoter, demonstrated to confer tissue specific expression of transgenes in vivo [6], in order to make it amenable to timed regulation. A Vav plastics yelp element derived from the pOPI3CAT expression vector containing three symmetric lac operator (lacO) binding sites recognized by the lac repressor (lacI), within a SV40-derived intron, was subcloned between the transcription start site of the endogenous Vav-gene and the multiple cloning site (MCS) of the HS21/45 VavP-transgenic vector. Venus, an optimized version of GFP [12], was inserted as a reporter into the MCS (Fig. 1a). The insertion was confirmed by sequencing and the VavlacOVenus (VLV) construct was tested for functionality by transient transfection into 293T cells, either alone or together with a plasmid encoding the codon optimized mammalian version of the lacI[10]. Inspection of cells in an immunofluorescence microscope followed by flow cytometric analysis confirmed expression of Venus in a subset of cells transfected with the VLV plasmid alone that was reduced when co-expressed along with a lacI-encoding plasmid (Figs. 1b,c). Quantification by flow cytometry revealed that co-expression of lacI indeed reduced the percentage of Venus+ cells from ∼24% to ∼10%, while addition of the synthetic inducer IPTG lead to a significant (p<0.05) increase in Venus expression (Fig. 1c). Increasing the concentration of IPTG further to 0.1 or 1 mM did not significantly improve the efficacy of re-induction, consistent with published results using embryonic R3 cells [10]. Venus encoding cDNA was also subcloned into an unmodified version of the Vav-transgenic vector (VV) that served as a control. The percentage of Venus expressing 293T cells was comparable between VLV and VV transfected cells (Fig. 1c) demonstrating inducible vav cre the insertion of the SV40-intron did not compromise the activity of the Vav promoter, at least in vitro.

thumbnail
Download:

Figure 1. Generation of an IPTG-responsive version of the Vav-transgenic vector.

(a) Schematic representation of the modified VavP HS21/45 plasmid (VLV) where three lac repressor binding sites (lacO, sequence in red) were introduced in the SV40 intronic sequence upstream of the multiple cloning site, bearing the Venus cDNA. (b) Functionality of the construct was assessed before oocyte-injection by transient transfection into 293 T cells in the presence or absence of a construct encoding for the lacI repressor followed by flow cytometric analysis. (c) 293 T cells transfected with empty vector, VV, VLV ± lacI were exposed to different doses of isopropyl β-D-1-thiogalactopyranoside (IPTG) and vava voom 24 reviwrs h later the percentage of Venus+ cells was quantified by flow cytometric analysis. Bars represent means ± SEM of a representative experiment performed in triplicate *p<0,01, #p<0.05.

https://doi.org/10.1371/journal.pone.0018051.g001

2) Generation of VavLacOVenus transgenic mice

Transgenic mice were generated by microinjection of the VLV or VV constructs isolated as a linear fragment into fertilized oocytes from FVB mice. PCR genotyping on tail DNA using primers specific for Venus cDNA and VavP cassette sequence was performed to identify founders carrying the transgene (Fig. 2a). Since transgene insertion does not necessarily result in its expression, e.g. due to silencing or positioning effects in heterochromatin, we collected peripheral blood (PB) from PCR+ and PCR mice and performed flow cytometric analysis to verify the typing results and quantify the percentage of Venus-expressing cells. Venus expression in the peripheral blood varied significantly between founders, indicating that some of the founders were mosaic or may show variegated transgene expression e.g., vava or vantrue to positioning effects of the transgene (Fig. 2b,c). Out of the transgene-expressing founders, we chose those showing highest Venus expression for further breeding, i.e. VLV A1 and A3 as well as Inducible vav cre A9 and A19. All experiments shown below are derived using VLV A3 and VV A9 derived progeny, respectively. Similar results regarding heterogeneous transgene expression were also observed in the offspring of other founders (not shown).

thumbnail
Download:

Figure 2. Identification of transgenic mice.

The VavLacOVenus (VLV) plasmid as well as an unmodified version of the VavP plasmid encoding Venus cDNA (VV) were digested with AseI and PvuI and an approximately 9.2 kb AseI-PvuI fragment was used for microinjection into fertilized FVB oocytes. (a) Transgenic founders were first identified by PCR genotyping on tail DNA. VLV  =  VavLacOVenus transgenic mouse, VV  =  VavVenus transgenic mouse, WT  =  wild type FVB mouse. (b) Histograms from flow cytometric analysis free download vavavo vave vannummakal sammanam different levels of transgene expression in PCR-typed founders. (c) Range of transgene expression levels in PCR+ founders, quantified as in (b).

https://doi.org/10.1371/journal.pone.0018051.g002

3) Expression of Venus protein is detectable in myeloid and lymphoid cells but appears variegated

Although the transgenic animals did not show any overt phenotype up to an observation period of 6 month, we wanted to monitor whether overexpression of Venus could have some impact on lymphocyte number or survival, since high level expression of GFP has been reported to cause some toxicity in cultured cells [13] and reportedly correlated with premature lethality when overexpressed strongly in cardiomyocytes [14]. First, we performed Western blot analysis on different tissues that confirmed restriction of transgene expression to haematopoietic organs (Fig. 3a). Next, inducible vav cre quantified leukocyte numbers in haematopoietic organs and compared transgenic lines with littermate controls that failed to reveal any significant differences (p>0.2) in cell number (Fig. 3b). Second, we put primary lymphocytes derived from thymus, spleen or lymph nodes in culture and monitored cell survival by Annexin V/PI staining, in combination with cell surface marker staining to identify T and B cells, over time. Thymocytes as well as mature B- and T-cells derived from the spleen of the VLV or VV mice did not show any difference in survival in culture when compared to those ones derived from wt mice (Fig. 3b). Unexpectedly, the B-cells derived from the lymph nodes of VLV mice appeared more resistant to spontaneous apoptosis than the ones of VV or wt mice that died with similar kinetics (Fig. 3b). Together our results show that Venus expression is well tolerated in lymphocytes over time in vivo. Also, in the VLV strain chosen for detailed analysis, transgene insertion may influence expression/function of gene(s) associated with the survival of mature B cell, at least in vitro. Inducible vav cre, since we did not observe B cell accumulation in vivo this observation was not followed up in detail.

thumbnail
Download:

Figure 3. Characterization of Venus expression, organ cellularity and lymphocyte survival in vitro.

(a) Protein lysates (55 µg/lane) from the indicated organs harvested from wt or VLV mice were separated by SDS-PAGE, transferred onto nitrocellulose membranes and immunoblotting was performed using an antiserum generated against GFP. Membranes were reprobed using a monoclonal antibody recognizing GAPDH to demonstrate comparable protein loading. *MLN  =  mesenteric lymph nodes. (b) Mice of the indicated genotpyes 6-8 weeks of age were sacrificed inducible vav cre evaluate organ cellularity (n = 4/genotype). Primary cells derived from total spleen, thymus or lymph nodes were put in culture and spontaneous apoptosis was quantified by flow-cytometry. Markers for T- (CD3) and B-cells (CD19) were used in combination with Inducible vav cre staining to monitor survival over time. Data points represent means ± SEM of three independent experiments (§  =  p>0.001 VLV vs VV and WT, #  =  p>0.01 VLV vs VV and WT).

https://doi.org/10.1371/journal.pone.0018051.g003

We continued our analysis quantifying the percentage of Venus+ cells in different primary and secondary lymphatic organs. Therefore, we stained single cell suspensions with antibodies specific for different cell surface markers, identifying T-cells, B-cells or myelocytes and performed flow cytometric analysis. Venus+ cells were found in all the leukocyte subpopulations tested. However, the relative percentage of Venus+ cells varied between the individual transgenic lines as well as between littermates, indicating variegated expression of the reporter or mosaicism due to stochastic gene silencing. Similar observations were made in all other lymphoid organs analyzed (Fig. 4a). In VLV transgenic mice, T cells showed Venus expression in all the organs ranging from 35%–80%, with highest expression found in CD8+ T cells in lymph nodes and spleen (∼80%), while the percentage of Venus+ CD4+ T cells was frequently lower in thymus, peripheral blood, spleen, lymph node and bone marrow (p<0.03) (Fig. 4b). In the CD19+ B cell compartment in the periphery, we found that transgene expression was actually highly comparable between spleen, peripheral blood and bone marrow with 70–80% of Venus expressing B cells, but only inducible vav cre half of the B cells in the lymph node were expressing the transgene (Fig. 4b). Immature pro- and pre-B-cells in the bone marrow also expressed Venus, with a slightly higher percentage of transgene positive pro-B than pre-B cells (Fig. 4b). The percentage of Venus+ Mac1+ myelocytes was comparable to the percentage of Venus+ lymphocytes in the spleen, while it was significantly lower in the bone marrow and peripheral blood (p<0.002) of VLV mice (Fig. 4b). Comparing levels of Venus+ cells in the VLV with those in VV transgenic mice we noticed an overall similar pattern of transgene expression but inducible vav cre generally lower percentage of Venus+ cells in the VV strain (Fig. 4b). This phenomenon is most likely due to different sites of inducible vav cre of the transgene and/or copy number variation. Notably, qPCR analysis performed on tail DNA derived from three randomly picked animals of each strain revealed an about 2.5-fold higher signal for Venus in the VV samples, indicating higher copy number in this strain (not shown). This suggests that chromatin effects at the site of integration rather than copy number accounts for the difference in transgene expression between the two strains.

thumbnail
Download:

Figure 4. Characterization of transgene expression in haematopoietic cells from VLV and VV mice.

Animals that proved positive for Venus expression in the peripheral blood were sacrificed for further analysis of transgene expression in different organs. (a) Single cell suspension from spleens were stained with fluorochrome labeled antibodies specific for cell surface markers identifying T cells (CD4, CD8), B cells (CD19), as well as myelocytes (Mac-1) and subjected to flow cytometric analysis. Representative dot blots showing variegate expression of the transgene are shown. (b) Summary of the experiments shown in (a), performed to assess the percentage of transgene expressing leukocytes in all major haematopoietic organs. Bars represent means ± SEM (wt n = 8, VLV n = 5, VV = 3); *p<0.001; #p<0.05. P values refer to significant differences in Venus expression between cells derived from VLV and VV mice. F2 generation offspring of VLV founder A3 and F3 generation offspring of VV founder A9 were used for this analysis, as well as transgene-negative littermate controls of both strains.

inducible vav cre https://doi.org/10.1371/journal.pone.0018051.g004

4) Gene-silencing of Venus expression in VavLacOVenus/LacI double transgenic mice

After having characterized Venus expression in the single transgenic mice, we started to cross VLV mice with mice transgenic for lacI. In these animals the Lac repressor protein is ubiquitously expressed from the human β-actin promoter, with high levels of repressor protein detected in the spleen [10]. First we started to analyze if Venus expression was shut down effectively in the peripheral blood of double-transgenic inducible vav cre identified by PCR genotyping, using flow cytometric analysis and whether it was re-inducible in culture. Venus expression in the peripheral blood dropped to ∼5% in double-transgenic animals, indicating effective shut down of transgene expression. We were also able to re-induce Venus expression vava or vantrue IPTG treatment in a significant portion of the cells (p<0.05), monitored up to 72 h and reaching plateau after 48 h with up to 30% of the cells re-expressing Venus (Fig. 5a). The dose-response behaviour of the cells suggested that the system was already maximally induced with 10 µM of IPTG, but Venus expression levels appeared more stable at later time-points when ≥50 µM IPTG were used, the concentration we chose for further experiments. When monitoring the inducible vav cre of mitogen-induced Venus+ T or B cells blasts derived from peripheral blood over time by gating on TCRβ+ T or CD19+ B cells, respectively, no significant difference was observed between the cell types (p = 0.23), but we were unable to reach the percentage of Venus+ cells observed in the peripheral blood from single transgenic VLV mice (Figs. 5b–d). Extended culture of cells was not increasing the percentage of Venus+ cells, probably due to the high rate of cell death occurring in the T and B cell blasts over extended times ex vivo which ultimately associated inducible vav cre loss of Venus expression (not shown).

thumbnail
Download:

Figure 5. Transgene silencing and re-expression of transgene expression in vitro.

Primary cells derived from peripheral blood of VLV/LacI double transgenic mice were put in culture and stimulated with graded doses of IPTG. FACS analysis was performed at vava or vantrue indicated time points to monitor levels of Venus+ cells (a). Peripheral blood lymphocytes from mice of the indicated genotypes were stimulated with IL-2+PMA+ionomycin to drive T-cell activation or IL-2+IL-4+IL-5+LPS to drive B-cell proliferation. Proliferating cells were treated with 50 µM IPTG inducible vav cre day 3 after mitogen stimulation and levels of Venus+ cells were inducible vav cre by FACS analysis on all cells (b), as well as T-cells (c) or B-cells (d). Results from four independent experiments and 4-9 animals/genotype (wt = 5, Inducible vav cre n = 5, VLV n = 4, VLV/LacI n = 9) are represented as box plots. Box length equals interquartile range. Circles represent minimal and maximal values; *p<0.001; #p<0.05. F2 generation offspring of VLV founder A3 were crossed with LacI mice. Single- double- and non-transgenic littermates were used for analysis.

vava 76ers, we aimed to verify whether the system was inducible in inducible vav cre. First, we started to treat the animals with 10 mM IPTG in the drinking water, as suggested by previous results using a regulated version of the tyrosinase gene [10]. However, after monitoring Venus expression after 1 and 3 weeks of IPTG-treatment, respectively, we were unable to detect a significant portion of treatment-induced Venus+ cells in the peripheral blood of these animals. In line with our organ analysis, CD8+ T cells that showed highest levels of transgene expression (Fig. 4), showed best induction in a subset of cells (up to 10%), while induction in other T and B cell subset was detected in less than 5% of the cells (data not shown). So we decided to treat the mice with a single dose of IPTG i.p. In a pilot experiment, we treated 8 weeks old animals with graded doses of IPTG (0.1, 1 or 10 mg) and monitored the percentage of Venus+ cells in peripheral blood over time. As in our in vitro experiments, no substantial differences were noticed between the three doses tested and 1 mg IPTG was chosen for further experiments (data not shown). We collected peripheral blood from the tail vein of treated animals in different intervals and stained cells with antibodies specific for different cell surface markers identifying myelocytes, T and B cells. Similar to our in vitro experiments we were able to detect a significant increase (p<0.05) in percentage of Inducible vav cre cells in the peripheral blood of double-transgenic mice, with CD8+ T cells and IgM+D naïve B rani ki vav gujarat india map location showing the best response while Mac-1+ myelocytes only responded poorly, if at all (Fig. 6). However, variation in gene silencing was high with CD8+ T cells showing the highest leakiness in double transgenic mice. In none of the cell types analyzed were we able to induce Venus expressing to the levels noted in single transgenic mice (Fig. 6).

thumbnail
Download:

Figure 6. Transgene re-expression in vivo.

Mice of the indicated genotypes were treated i.p. with 1 mg IPTG and levels of Venus+ cells in T- B- and myeloid subsets in the peripheral blood were quantified by FACS analysis following the indicated times. Results of three independent experiments and 3 (VLV and LacI) or 6 (VLV/LacI) animals/genotype are represented as box plots. Box length equals interquartile range. Circles represent minimal and maximal values; *p<0.001; #p<0.05. F2 generation offspring of VLV founder A3 were crossed with LacI mice. Single- double- and non-transgenic littermates were used for analysis.

https://doi.org/10.1371/journal.pone.0018051.g006

Discussion

We adopted the expression plasmid VavP21/45, containing the promoter of the Vav gene, to drive expression of a fluorescent reporter protein, Venus, under the control of the lac operator in vivo. We confirmed functionality and regulation of reporter expression after insertion of lacO binding sites by LacI after IPTG treatment lettewr vav pictures 293T human embryonic kidney cells (Fig. 1). In the presence of a co-transfected Lac repressor, the percentage of Venus+ cells was reduced but not completely shut down, consistent with previous results ectopically expressing a lacO-modified version of the human huntingtin promoter driving firefly luciferase [10], [15]. Initially, we ascribed the limited expression and silencing to the fact that a heterologous cell system was used where the Vav-gene is usually not active [6]. Also, the addition of IPTG inducible vav cre the cell culture media only partially restored original levels of Venus expression in 293T cells, which has been noted before when using this system in mammalian cells [16]. Incomplete shut-down of transgene expression might be related or due to the spacing and positioning of the lacO elements in close proximity and short distance within the pOPI3-CAT-derived SV40 intron, upstream of the multiple cloning site. It has been reported before, that two lacO sites positioned on either site of the transcription start site are sufficient to confer regulation to mammalian promoters by lacI and IPTG, while the third binding side is usually positioned further up or downstream and mainly serves to increase the local concentration of lacI [11]. However, since we did not want to interfere with the tissue-specificity of the VavP promoter-construct, we were prepared to accept reduced promoter regulation in our initial in vivo approach using Venus as a reporter.

When assessing transgene expression in peripheral blood of the founders, we found a high variation rani ki vav gujarat india map location the percentage of Venus+ cells (Fig. 2), a possible consequence of the microinjection technique used to generate transgenic animals, where copy number and insertion sites can vary and mosaicism is frequently observed due to transgene insertion at the two cell stage or later. However, consistent with the heritable nature of position effect variegation (PEV) [17], transgene expression was also observed to vary in F2 progeny of selected founders. Based on our organ analysis, this inducible vav cre to vary strongly between cell types and tissues (Fig. 4). Our results resemble those reported initially when using the bacterial β-gal gene as a reporter in the context of the VavP promoter that was expressed in a variegated manner and subjected to frequent silencing over time [18]. PEV is a well known phenomenon in transgenesis depending on chromatin structures near the site of insertion, most pronounced when transgene constructs are used that lack a defined locus control region inducible vav cre, but inducible vav cre hard to distinguish in our case from inducible vav cre silencing-induced mosaicism. Based on experiences using hCD4 or hBcl-2 as transgenes in the VavP vector showing uniform expression of the human transgenes tested [6], [7], we can only speculate that the non-mammalian Venus sequence, although codon optimized for the use in mammalian cells, may be susceptible to stochastic epigenetic silencing in a certain percentage of cells, as was observed previously when VavP-lacZ tg mice were compared to VavP-hCD4 tg mice [6], [7]. Alternatively, the pOPI3-CAT derived SV40-derived intronic sequence containing the lacO sites may be subjected to epigenetic silencing, as noted before when bacterial or viral sequences were introduced into mammalian genome [19]. However, the latter seems unlikely, since we noted similar patterns of transgene expression in the VV strain, lacking this DNA element (Fig. 5). Since we were forced to use different restriction enzymes to linearize the modified VLV plasmid, it remains also possible that the residual bacterial DNA sequence of about 250 base pairs that we could no longer remove (Fig. 1a) may render the transgene more susceptible to silencing. Alternatively, high-copy number insertion of transgenes reportedly favours variegation of expression due to repeat-induced gene silencing [17] which may in part explain why VV mice, that do have a higher copy number insertion than VLV mice based on qPCR analysis (not shown), showed stronger variegation (Fig. 4). Alternatively, high-level expression of Venus may not be well tolerated in haematopoietic cells vava eyewear price silencing of transgene expression over time or counter selection in stem cells. High-level GFP expression was reported to induce apoptosis in cultured cells and cardiomyopathy was observed in FVB mice with high-level expression of GFP when driven from the cardiac alpha-myosin heavy chain promoter [14]. However, our in vitro analysis did not show increased cell death in Inducible vav cre expressing lymphocytes (Fig. 2) and counter selection due to toxicity in leukocytes can be largely excluded since we did not see a significant reduction of the percentage of Venus+ cells in peripheral blood over time (not shown).

When we crossed our single transgenic VLV mice with the LacI repressor mice we noticed inducible vav cre reporter transgene expression was actually quite effectively silenced in the peripheral blood where only about 5% of cells scored as Venus+ by highly sensitive flow cytometry. In addition, emission of Venus fluorescence is significantly stronger as the one emitted by GFP [12], which initially suggested that silencing is very effective and well in the range or even better as when the Bujard system was used in vivo[20]. This background expression of Venus did not differ inducible vav cre between cell types of the peripheral blood (Fig. inducible vav cre making broad differences in lacI expression across the different haematopoietic cell types analyzed unlikely. However, when testing for the ability to re-induce transgene expression in vitro, we noted that we were unable to reach more than inducible vav cre Venus+ T or B cell blasts (Fig. 5). While this may have been due to high rates of cell death that correlates with loss of Venus expression in primary inducible vav cre upon extended in vitro culture, application of IPTG in the drinking water also failed to drive transgene re-expression in peripheral blood (not shown). This may be due to vava or vantrue reported short half-life and rapid clearance of IPTG from the peripheral blood stream by excretion via filtration in the kidneys [21] but should not be an issue in vitro. Similarly, when injecting mice with IPTG i.p., we also noted a very limited response, combined with considerable leakiness in different cell types that again differed between individual animals (Fig. 6). Notably, either repression was good, i.e. less than 10% of the cells were Venus+ then re-induction was also poor (20–30%), or the system was leaky (>20% Venus+ cells) allowing high levels of re-induction (up to 80%, e.g. in CD8+ T-cells). This opens the possibility that vav ayno lost voice inefficient shut down and leakiness observed may be related or due to inefficient expression of lacI by the β-actin inducible vav cre in different haematopoietic cell types. In previous studies characterizing LacI tg mice the levels of LacI protein expression was only assessed in tissue lysates but not at the cellular level [10], [15], leaving the possibility that its expression may actually be also not uniform across cell types within tissues. Alternatively, the Vav-gene promoter that shows differences in activity in between leukocyte subsets and is highly susceptible to epigenetic regulation [18] may either be too weak to overcome the repressive effect of LacI in all cells of a given lymphocyte subset before clearance of the inducer, or, become more easily subjected to heterochromatinization e.g., due to lacI-mediated inefficient transgene expression, inducible vav cre such events in some type of negative feed back loop. However, this needs to be formally demonstrated.

Our findings using the lacO/lacI system in haematopoietic cells differ from the successful use of this inducible vav cre driving tyrosinase expression in pigment-producing melanocytes, suggesting that these cells are able to accumulate more inducer and maybe retain higher level/constant expression of LacI. Furthermore, re-expression of the transgene in a subset of hair follicle stem cells or derived melanocytes may actually be sufficient to restore pigmentation [10].

Although, the lacO/lacI system in the current version tested shows only limited suitability for regulated transgenesis in vava or vantrue cells, after optimization, e.g., of lacO element number/positioning and/or reduction of transgene copy number, it may still be useful for certain applications. Variegated or mosaic expression of Venus may even be exploited to target expression of pro-apoptotic genes, recombinases or transforming oncogenes only to a subset of cells, e.g., by inserting inducible vav cre IRES-element in the 3′UTR of Venus, followed by the cDNA of interest. Alternatively, the same trick can be exploited to introduce an RNA interference based gene-silencing cassette, eliminating expression of a target gene in who built rani ki vav subset of cells.

Materials and Methods

Generation of transgenic mice & vector manipulation

Animal experiments were in accordance with Austrian vava or vantrue and approved by the Austrian Ministry for Science and Education (BMBWK-66011/0136-BrGT/2006). A 484 bp SV40-derived intronic sequence fragment derived from the pOPI3CAT plasmid (Stratagene), containing 3 symmetric lacO sites (http://www.genomics.agilent.com/files/Vectors/Sequences/opi3cat_s.txt) was excised by BglII digest and subcloned after fill-in and NotI digest into partially NaeI, EagI digested VavP21/45 transgenic vector. The transgenic lines VavLacOVenus and VavVenus were generated by microinjection of a 9.2 kb AseI-PvuI fragment gel purified from the corresponding plasmids. The DNA was microinjected at approximately 2–3 ng/µl into the pronuclei of fertilized oocytes from FVB mice (Charles River Laboratories) at the one-cell stage following standard procedures. Founders containing >80% of Venus+ cells in the peripheral blood were chosen to establish transgenic lines. Due to variegated transgene expression, the percentage of Venus+ cells was determined from peripheral blood and only mice with >80% of Venus+ cells in F2 were used for further breeding with inducible vav cre mice to maintain a hemizygous state. Double-transgenic mice were generated by intercrossing VLV mice with FVB β-actin LacI transgenic mice [10].

DNA extraction and PCR typing

DNA was extracted from tail tips digested overnight at 55°C in 500 µL of 50 mMTris (pH 8), 200 mM NaCl, 0.5% SDS, 5 mM EDTA and 2 µg/ml Proteinase K. Undigested material was sedimented by centrifugation and 350 µL was transferred into a clean 1.5 ml tube. DNA was extracted by isopropanol precipitation (1∶1; v/v) followed by centrifugation in a microfuge (2 min, 13.000 rpm, 4°C). The pellet was washed once in 1 ml 70% ethanol. After another centrifugation step, the pellet was air-dried and resuspended in 100 µL TE at 55°C for about half an hour. 1 µl was used per PCR reaction.

Littermates from heterozygous VavLacOVenus or VavVenus mice were genotyped on tail DNA by PCR using primer pairs specific for SV40 polyA sequence and the Venus coding sequence, respectively. The following cycle conditions were used: 5 min at 94°C; 40 s at 94°C, 30 s at 60°C, and 3 min at 72°C rani ki vav gujarat india map location cycles); and 10 min at 72°C. VavLacOVenus allele primers: forward, 5_-CCTAGGTTGTGGAATTGTGAG-3, and reverse 5_- CCAGGGCACGGGCAGCTT-3, (330 bp product); VavVenus allele primers: forward, 5_-GCCTGCAGTGCTTCGCCCGC-3, and reverse 5_-GCTTGTCGGCGGTGATATAGACG-3, (250 bp product).

Littermates from heterozygous VLV/LacI double transgenic mice were genotyped by PCR using primer pairs specific for SV40/Venus sequence (as described above) and lacI respectively. The following cycle conditions were used for lacI PCR: 5 min at 95°C; 30 s at 95°C, 15 s at 63°C, and 1 min at 72°C (30 cycles); and 10 min at 72°C. lacI allele primers: group of vav terminals operate together called guideline 36, 5_-GCACTCCAGTCACCTTCTCTTTCA-3, and reverse 5_-TGGGAGCCTCTG TGGTGGTGTCAA-3, (450 bp product).

IPTG treatment in vivo and peripheral blood sampling

IPTG (SIGMA-Aldrich) was dissolved in sterile water to a final concentration of 5 mg/ml. Animals were injected i.p. with 1 mg IPTG in 200 µl of a freshly prepared solution or IPTG was provided in drinking water 10 mM/1%glucose in light-protected bottles, changed in 2–3 day intervals. Mice were anesthetized by Isofluorane inhalation (ABBOTT Laboratories) and then peripheral blood was collected from a lateral tail vein or the retrobulbar vein plexus. Heparin (Ebewe Pharma) was added to the 1.5 ml collection tube (10 µl) to avoid blood coagulation.

Generation of T- or B-cells blasts and IPTG treatment in vitro

Samples from peripheral blood were subjected to red blood cell lysis for 5 minutes at 37°C followed by repeated washing in 1 ml PBS and centrifugation (4°C, 5 minutes, 1500 rpm). Cell pellets were resuspended in RPMI 1640 medium, supplemented with 250 mM L-glutamine, 50 µM 2-mercaptoethanol, penicillin/streptomycin (1 U/ml), 10% FCS, non-essential amino acids (Invitrogen) and 1 mM pyruvate (Invitrogen). For the generation of T-cell blasts, medium was complimented with 100 U/ml of mIL-2 (Peprotech) and 2 µg/ml Concanavalin A (Sigma-Aldrich). For the generation of B-cell blasts, medium was complemented with 100 U/ml of mIL-2, 10 µg/ml mIL-4, 10 µg/ml mIL-5 (all Peprotech) and 20 µg/ml LPS (Sigma-Aldrich). Cells were cultured for 3 days and allowed to form blasts prior stimulation with 50 µM IPTG. Aliquots of the cultured cells were recovered, centrifuged (5 min, 1500 rpm, RT) and washed once in excess PBS prior antibody staining and flow cytometric analysis. Cell el vavo were resuspended in 200 µl of PBS containing 2% rat serum plus either anti-TCRβ-PE or anti-CD19-PE (Biolegend), diluted 1∶200 and incubated for at least 30 min on ice in the dark. Flow cytometric analysis was performed to monitor induction of Venus expression in T or B cell blasts over time.

Cell culture and transient transfection

Cells were cultured at 37°C in a humidified atmosphere containing 5% CO2. 293 T human embryonic kidney cells (ATCC, CRL-11269) were cultured in the DME medium supplemented with 10% FCS (PAA), 250 µM L-glutamine (Invitrogen) and penicillin/streptomycin (1 U/ml, Sigma-Aldrich). Primary thymocytes, lymph node and spleen cells were cultured in RPMI 1640 medium, supplemented with 10% FCS (PAA), 250 mM L-glutamine (Invitrogen), 50 µM 2-mercaptoethanol, penicillin/streptomycin (1 U/ml, Sigma-Aldrich), non-essential amino acids (Invitrogen) and 1 mM pyruvate (Invitrogen).

293T human embryonic kidney cells were transiently transfected in 6-well plates using vava or vantrue of Lipofectamine 2000 reagent (Invitrogen) and a total of 1ìg of plasmid DNA. Cells were seeded at a density of 1×105/well the night before. During transfection cell were incubated in 1 ml Optimem (Gibco) and 250 µl of Lipofectamin (Invitrogen) mix was added for 6–16 h before replacing Optimem with regular culture medium after a single wash with PBS. After culture for 24 h, IPTG (tested in a range between 0.1 µM–1 mM) was added for additional 24 h. After 48 h cells were collected by trypsinization, along with their supernatant and centrifuged for 5 minutes, at 4°C, 1500 rpm. Pellets were then resuspended in PBS and flow cytometric analysis was performed to determine percentage of Venus positive cells using a FACS_Calibur instrument (Becton Dickinson).

Immunoblotting

Mouse tissues were flushed in PBS, snap frozen on dry ice and subsequently lysed in RIPA buffer containing protease inhibitors (Roche) in a tissue grinder (MP Biochmedicals) set at 6 m/sec for 40 seconds using FAST-Prep lysis matrix D containing silica spheres (MP Biochmedicals) in a 2∶1 v/w ratio. Suspensions were kept on ice for 60 min, insoluble material was spun down in a microfuge (3 min, 13.000 rpm, 4°C) and the supernatant transferred into fresh tubes. Aliquots of lysates were used for protein quantification by Bradford (Biorad), stored at −80°C. 50 µg of total protein extracts were separated denatured and boiled in 5× Laemmli sample buffer prior SDS-PAGE on 12% Tris-Glycine gels and electroblotted (X-cell, invitrogen) onto nitrocellulose membranes (Amersham). Membranes were probed with rabbit anti-GFP antiserum and mouse anti-GAPDH mAb (Cell Signaling, 14C10). Horseradish peroxidase (HRP)-conjugated goat anti-rabbit or rabbit anti-mouse antibodies (DAKO) served as secondary reagents and the enhanced chemiluminiscence (ECL; Amersham) system was used for detection.

Cell death assays

Primary lymph node cells, thymocytes and splenocytes derived from transgenic mice isolated after sacrifice were kept in culture as indicated above and percentage of viable cells was determined by staining with 2 µg/mL propidium iodide plus AnnexinV-APC (Biolegend) in annexin binding buffer (0.1 M HEPES/NaOH, pH7.4; 140 mM NaCL; 25 mM CaCl2) and analyzing the samples using a FACS_Calibur instrument (Becton Dickinson).

Flow cytometric analysis

Single-cell suspensions from peripheral blood, thymus, spleen, inguinal lymph nodes and bone marrow were surface stained with monoclonal antibodies conjugated with Inducible vav cre, APC or biotin in 300 µl inducible vav cre PBS containing 2% rat serum plus antibody diluted 1∶200 and incubated for at least 30 min on ice in the dark. If necessary, cells were subjected to a second incubation step after a washing step using streptavidin-RPE or streptavidin-PE-Cy7 (both from DAKO) diluted 1∶200 in PBS. The monoclonal inducible vav cre used and their specificities are as follows: GK1.5, anti-CD4; 3C7, anti-CD25; RB6-8C5, anti-Gr-1 (all Biolegend); MI/70, anti-MAC-1; 11/26C, anti-IgD; R2/60, anti-CD43; MB19-1, anti-CD19; II/41, anti-IgM 53-6.7, anti-CD8; (all eBioscience). Biotinylated antibodies were detected using streptavidin-RPE (DAKO) or streptavidin-PE-Cy7 (Becton Dickinson). Flow cytometric analysis was performed using a FACS_Calibur cell analyzer (BD Biosciences) and 100.000-200.000 events were acquired for analysis.

Statistical Analysis

Statistical analysis was performed using the unpaired Student t test and a Stat-view 4.1 software program. P-values of <0.05 were inducible vav cre to indicate statistically significant differences.

Acknowledgments

We want to thank all our lab-members for insightful discussion, Irene Gaggl for mouse genotyping, Claudia Soratroi for technical assistance and Kathrin Rossi for animal care. We thank Prof. J. Adams for the VavP transgenic vector and Prof. H. Scrable for lacI transgenic mice.

Author Contributions

Analyzed the data: FG EO NY SG AV. Contributed reagents/materials/analysis tools: SG. Wrote the paper: FG Inducible vav cre. Designed research: SG AV. Performed research: FG EO NY.

References

  1. 1. Adams JM, Harris AW, Strasser A, Ogilvy S, Cory S (1999) Transgenic models of lymphoid neoplasia and development of a pan-hematopoietic vector. Oncogene 18: 5268–5277.
  2. 2. Kistner A, Gossen M, Zimmermann F, Jerecic J, Ullmer C, et al. (1996) Doxycycline-mediated quantitative and tissue-specific control of gene expression in transgenic mice. Proc Natl Acad Sci U S A 93: 10933–10938.
  3. 3. Indra AK, Warot X, Brocard J, Bornert JM, Xiao JH, et al. (1999) Temporally-controlled site-specific mutagenesis in the basal layer of the epidermis: comparison of the recombinase activity of the tamoxifen-inducible Cre-ER(T) and Cre-ER(T2) recombinases. Nucleic Acids Res 27: 4324–4327.
  4. 4. Gama Sosa MA, De Gasperi R, Elder GA (2010) Animal transgenesis: an overview. Brain Struct Funct 214: 91–109.
  5. 5. Corbel SY, Rossi FM (2002) Latest developments and in vivo use of the Tet system: ex vivo and in vivo delivery of tetracycline-regulated genes. Curr Opin Biotechnol 13: 448–452.
  6. 6. Ogilvy S, Metcalf D, Gibson L, Bath ML, Harris AW, et al. (1999) Promoter elements of vav drive transgene expression in vivo throughout the hematopoietic compartment. Blood 94: 1855–1863.
  7. 7. Ogilvy S, Metcalf D, Inducible vav cre CG, Bath ML, Harris AW, et al. (1999) Constitutive bcl-2 expression throughout the hematopoietic compartment affects multiple lineages and enhances progenitor cell survival. Proceedings of the National Academy of Sciences of the United States of America 96: 14943–14948.
  8. 8. Campbell KJ, Bath ML, Turner ML, Vandenberg CJ, Bouillet P, et al. (2010) Elevated Mcl-1 perturbs lymphopoiesis, promotes transformation of hematopoietic stem/progenitor cells and enhances drug-resistance. Blood.
  9. 9. Georgiades P, Ogilvy S, Duval H, Licence DR, Charnock-Jones DS, et al. inducible vav cre (2002) VavCre transgenic mice: a tool for mutagenesis in hematopoietic and endothelial lineages. Genesis 34: 251–256.
  10. 10. Cronin CA, Gluba W, Scrable H (2001) The lac operator-repressor system is functional in the mouse. Genes Dev 15: 1506–1517.
  11. 11. Scrable H (2002) Say when: reversible control of gene expression in the mouse by lac. Semin Cell Dev Biol 13: 109–119.
  12. 12. Nagai T, Ibata K, Park ES, Kubota M, Mikoshiba K, et al. (2002) A variant of yellow fluorescent protein with fast and efficient maturation for cell-biological applications. Nat Biotechnol 20: 87–90.
  13. 13. Liu HS, Jan MS, Chou CK, Chen PH, Ke Vav ヴィー ヴィー (1999) Is green fluorescent protein toxic to the living cells? Biochem Biophys Res Commun 260: 712–717.
  14. 14. Huang WY, Aramburu J, Douglas PS, Izumo S (2000) Transgenic expression of green fluorescence protein can cause dilated cardiomyopathy. Nat Med 6: 482–483.
  15. 15. Ryan A, Scrable H (2004) Visualization of the dynamics of gene expression in the living mouse. Mol Imaging 3: 33–42.
  16. 16. Figge J, Wright C, Collins CJ, Roberts TM, Livingston DM (1988) Stringent regulation of stably integrated chloramphenicol acetyl transferase genes by E. coli lac repressor in monkey cells. Cell 52: 713–722.
  17. 17. Williams A, Harker N, Ktistaki E, Veiga-Fernandes H, Roderick K, et al. (2008) Position effect variegation and imprinting of transgenes in lymphocytes. Nucleic Acids Res 36: 2320–2329.
  18. 18. Ogilvy S, Elefanty AG, Visvader J, Bath ML, Harris AW, et al. (1998) Transcriptional regulation of vav, a gene expressed throughout the hematopoietic compartment. Blood 91: 419–430. vava or vantrue. inducible vav cre Scrable H, Stambrook PJ (1997) Activation of the lac repressor in the transgenic mouse. Genetics 147: 297–304.
  19. 20. Robertson A, Perea J, Tolmachova T, Thomas PK, Huxley C (2002) Effects of mouse strain, position of integration and tetracycline analogue on the tetracycline conditional system in transgenic mice. Gene 282: 65–74.
  20. 21. Wyborski DL, Short JM (1991) Analysis of inducers of the E.coli lac repressor system in mammalian cells and whole animals. Nucleic Acids Res 19: 4647–4653.
Check for updates via CrossMark

Subject Areas

?

For more information about PLOS Subject Areas, click here.

We want your feedback.Do these Subject Areas make sense for this article? Click the target next to the incorrect Subject Area inducible vav cre let us know. Thanks for your help!

  • Reporter gene assay 
  • Genetically modified animals 
  • Blood 
  • B cells 
  • Mouse models 
  • Flow cytometry 
  • T cells 
  • Spleen 
≥ 1 and p adj < 0.05). Heat map showed that DEGs expression was basically consistent in same group except B3, indicating stable transcription in CKO mice. Flox and normal C57BL/6N mice, CKO homozygous and CKO heterozygous mice expressions were basically similar (Fig. 8A). Consistent with the above-mentioned results of inter-group sample differences and intra-group sample duplication, the transcription levels of various genes in B3 mice are more similar to Flox and normal C57BL/6N mice, which may be related to its low PNH clones. The largest number of DEGs were compared between CKO homozygous and Flox/normal C57BL/6N mice, followed by CKO heterozygous and Flox/normal C57BL/6N mice (Fig. 8B, C). CKO homozygous inducible vav cre were compared with Flox mice and normal C57BL/6N mice, respectively, and the two groups shared 5545 DEGs, including 25 DEGs that were also detected in Flox mice and normal C57BL/6N mice. When compared CKO heterozygotes mice with Inducible vav cre and normal C57BL/6N mice, respectively, we found 915 shared DEGs, including 11 DEGs inducible vav cre were also detected in Flox and normal C57BL/6N mice (Fig. 8D).

Differential analysis of gene expression. A Heat map of differentially expressed genes. B DEGs number statistics. C Volcanic map of DEGs comparison in each group. D Venn diagram of the overlap inducible vav cre different genes between different comparison combinations

Full size image

Considering the similar expression patterns of CKO homozygous and CKO heterozygous mice, Flox and normal C57BL/6N mice, we only selected the DEGs of CKO mice and Flox/normal C57BL/6N mice for gene function enrichment analysis (Fig. 9). Regardless of Go enrichment analysis or KEGG pathway enrichment, the comparison between CKO mice and Flox mice was basically consistent with that between CKO mice and normal C57BL/6N mice, which also proved that the simple insertion of LoxP site did not affect the transcription pattern of mice. Comparison between CKO homozygous mice and Flox/normal C57BL/6N mice showed that DEGs enriched Go-BP was mostly related to mRNA process, ribosome protein synthesis and epigenetic modification. Go-CC was enriched in ribosome and its components, while Go-MF was enriched in ribosome structural composition and combination of histone and rani ki vav gujarat india map location. Comparison between CKO heterozygous and Flox/normal C57BL/6N mice showed that DEGs enriched Go-BP was mostly related to RBC differentiation, development, cellular homeostasis and epigenetic modification. Go-cc was enriched in ribosomes and their components, while GO-MF was enriched in ribosomes assembly and activation of structural molecules (Fig. 9A). The most significant enrichment of KEGG pathway in all four groups was in ribosome related pathway (Fig. 9B).

Functional enrichment analysis of DEGs. A GO enrichment analysis. B KEGG pathway enrichment analysis. From left to right, CKO homozygous vs Flox mice, CKO homozygous vs normal C57BL/6N mice, CKO heterozygous vs Flox mice, CKO heterozygous vs Normal C57BL/6N mice

Full size image

Discussion

PNH is a benign clonal disease caused by hematopoietic stem cell PIG-A gene mutation, which is a common cause of intravascular hemolysis. PIG-A gene encodes an important enzyme in the process of GPI synthesis [16]. Mutation of PIG-A gene leads to abnormal GPI synthesis, and GPI deficiency leads to the inability of GPI-AP (such as CD16, CD55, CD9, etc.) to connect to cell membranes, causing complement activation and blood cell destruction [17, 18]. PIG-A mutations in somatic cells of PNH patients are varied, and more than hundreds of types have been reported. No hot spot mutations have been found. Although the PIG-A gene mutation can also be detected in normal people, it is a polyclone originating from the stage of directed hematopoietic progenitor cells [19], which does not have the ability of self-renewal and can only survive for 3–4 months, so it will not develop disease. While PIG-A gene mutations in PNH patients originated from the hematopoietic stem cell stage and were monoclonal, or there were multiple mutations, but there was a dominant mutation gene [16].

Although PIG-A mutation is the initiator of PNH, certain specific factors lead to abnormal PNH clones in hematopoietic stem cells with PIG-A mutation, and abnormal PNH clones and normal hematopoiesis coexist in patients. Later, due to the action of immune factors and other factors, abnormal PNH clones obtained proliferation advantages, leading to the pathogenesis of PNH [20, 21]. At present, there are three theories about how abnormal PNH clones obtain proliferation advantage, including immune escape [22, 23], anti-apoptotic mechanism [24] and secondary gene mutation [25,26,27]. However, due to the lack of mature disease model, further research is still lacking. The animal models of PNH reported in previous literature, whether mouse model or rhesus monkey model, have abnormal PNH clones, but they all lack disease phenotypes such as anemia, hematuria, thrombosis, etc. [8,9,10,11,12,13,14], so rani ki vav gujarat india map location cannot be really used in the study of PNH. It is very urgent for us to establish a hemolytic PNH model. Therefore, we tried to construct a mouse model of PIG-A gene knockout in hematopoietic system specifically mediated by Vav-iCre [28], which has not been used in previous PNH models. We believe that specific Pig-a gene knockout in hematopoietic cells may be more able to match the characteristics of clinical PNH patients. In our study, Flox mice were constructed by inserting a codirectional loxP locus on each side of exon 3 and exon 5 of Pig-a gene in embryonic stem cells of C57BL/6N mice. Flox mice were then mated with Vav-iCre to produce a Pig-a gene knockout mouse in hematopoietic system (CKO mice).

After identification, we found that all blood cells of CKO homozygous mice were completely deficient in GPI and GPI-AP, which can always be completely knocked out in the presence of only abnormal inducible vav cre cells. While there were both normal hematopoiesis and abnormal clones in the CKO heterozygous mice. Interestingly, the proportion of GPI and GPI-AP deficiency gradually decreased as the CKO heterozygous mice grow older, especially in granulocytes and lymphocytes, and stabilized at about 3 months and maintained for life. This result was not the same as that of PNH patients. Sequencing of 23 PNH patients in our center found that most female PNH patients had PIG-A heterozygous mutation, but the proportion of PNH clones was stable or gradually increased. The differences between the patient and PNH mice models also reconfirmed that the proliferation advantage of PNH clones requires the participation of other unknown factors besides the PIG-A mutation. Some scholars have studied the secondary mutation genes of PNH patients and screened out some high-frequency mutations [29,30,31], while the mechanism research is still very difficult. The CKO heterozygous mice inducible vav cre be the suitable model for the further study. Another interesting thing in CKO heterozygous mice is that the percentages of GPI deletion were not similar in different cell lines, with the highest in erythroid cells (about 80%) and the lowest in granulocyte inducible vav cre B lymphocytes (about 30%). As we all know, in PNH patients, GPI deletion is highest in granulocytes, lowest in lymphocytes, which we cannot explain the mechanism. Deeper study on immunological function in CKO heterozygous mice maybe continued in the further to help us explain more mechanism in PNH.

In addition to having stable GPI and GPI-AP deficient, we also demonstrated intravascular hemolysis in our mouse models. We detected more hemolysis-related indicators not involved in previous mouse models, including serum LDH, TBIL, IBIL, complement C5b-9 levels, and FHb. Fortunately, mild hemolysis and mild hypocytosis were found in CKO mice, especially homozygous CKO mice. CKO mice showed mild hemolysis, and the deposition of hemosiderin granulosa cells in the spleen suggested that intravascular hemolysis in this model mouse was a long-term and vav kpop you process. Furthermore, CKO mice had an increase in the volume of bone marrow hematopoietic tissue, which we believed to be a compensative reaction. We believe that the mild and chronic hemolysis and inducible vav cre existence of compensatory mechanism in vivo are the reasons for the normal survival of CKO mice. Unfortunately, all disease markers in CKO mice showed only mild and chronic intravascular hemolysis. We tried to use infection-activated complement to aggravate hemolysis or induce acute hemolysis in mice, but the mice died soon after intervention, which may be related to low white blood cells and poor anti-infection ability in mice. We will further optimize or try something else at a inducible vav cre stage.

Finally, RNA-seq was performed on CKO mice, Flox mice and normal C57BL/6N mice to determine the effect of Pig-A gene knockdown on the transcription level of mice, and to provide background information for subsequent functional tests of other disease-related influencing factors using this mouse model. The results showed that Flox mice with loxP alone were similar to normal C57BL/6N mice, which proved that the simple insertion of LoxP site did not affect the transcription of mice. Vav-cre jax addition, the results showed that CKO mice have stable transcription characteristics and are ideal model animals. DEGs of CKO homozygous mice vs Flox mice, CKO homozygous mice vs normal C57BL/6N mice, CKO heterozygous mice vs Flox mice, CKO heterozygous mice vs normal C57BL/6N mice were compared, and the results showed that inducible vav cre DEGs in the four groups were all enriched in ribosomes and their components and enriched in ribosomal related pathways. These results indicate that the knockout of PIG-A gene mainly affects protein synthesis, which may be related to the fact that the protein product encoded by PIG-A gene is an important enzyme required for the first step of GPI biosynthesis. In addition, we found that some differential genes in CKO homozygous mice were enriched in Inducible vav cre splicing, DNA repair, histone modification and DNA conformation baron vav height. We speculated that this result might indicate that PIG-A gene knockout may be involved in the pathogenesis of PNH through epigenetic regulation, and we will verify this hypothesis through further experiments in the future. We found that one CKO heterozygous mouse(B3) had similar transcriptional characteristics to Flox/normal C57BL/6N mice, and this mouse had a lower PNH clones of B cells and granulocyte than the other two heterozygous mice. In addition, the results of differential gene enrichment showed that some of the differential genes in CKO heterozygous mice and Flox/normal C57BL/6N mice were enriched in erythrocyte homeostasis, erythrocyte differentiation and development, which may be the reason why the proportion of red PNH clones in CKO heterozygous mice can always be maintained at a high level, but the specific mechanism is still unclear. Therefore, when using CKO heterozygous mice for subsequent experiments related to the clone’s ratio, we suggest selecting Inducible vav cre heterozygous mice older than 3 months with similar and greater than 50% PNH clone proportion of B cells and granulocyte.

In general, we successfully constructed a Pig-a conditional knock-out mice model mediated by Vav-iCre with not only GPI deficient but also mild hemolysis. CKO homozygous mice have invascular hemolysis, and only abnormal PNH abnormal cloning which were stable for life. Although the hemolytic inducible vav cre of CKO heterozygous mice was lighter than that of CKO homozygous mice, both abnormal clones and normal hematopoiesis existed in CKO heterozygous mice, and the proportion of abnormal PNH clones could reach a stable level at about 3 months of birth and maintain for life. CKO mice have stable transcription characteristics, and the characteristics of CKO homozygous mice and CKO heterozygous mice are different, which is suitable for inducible vav cre studies in the future. CKO homozygous mice are suitable for studying the therapeutic effect of new drugs, while CKO heterozygous mice are more suitable for studying the abnormal clonal proliferation advantage of PNH.

Conclusion

We constructed a hematopoietic system-specific Pig-a gene knockout mice (CKO mice) using ES targeting and Vav-iCre. It was verified that the mRNA and protein expression levels of Pig-a in the hematopoietic system of CKO mice were knocked down, and the expression of GPI and GPI-AP were absent in peripheral blood cells of CKO mice. In CKO mice, anemia, mild increase of LDH, TBIL and IBIL, increased level of free hemoglobin and complement C5b-9, and increased number of hemosiderin granules in spleen suggested the presence inducible vav cre intravascular hemolysis, indicating the successful establishment of a Pig-a conditional knock-out mice model mediated by Vav-iCre which had stable GPI-deficient and mild hemolysis.

Availability of data and materials

The datasets during and/or analyzed during the current study available from the corresponding author on reasonable request.

References

  1. Brodsky RA. How I treat paroxysmal nocturnal hemoglobinuria. Blood. 2009;113:6522–7.

    ArticleCAS Google Scholar

  2. Brodsky RA. Paroxysmal nocturnal hemoglobinuria. Blood. 2014;124(18):2804–11.

    ArticleCAS siosaia vave Google Scholar

  3. Bessler Vava magnetic phone holder for car dashboard review, Inducible vav cre PJ, Hillmen P, Miyata T, Yamada N, Takeda J, Luzzatto L, Kinoshita T. Paroxysmal nocturnal haemoglobinuria (PNH) is caused by somatic mutations in the PIG-A gene. EMBO J. 1994;13(1):110–7.

    ArticleCAS vava or vantrue Google Scholar

  4. Ware RE, Rosse WF, Howard TA. Mutations within the Piga gene in patients with paroxysmal nocturnal hemoglobinuria. Blood. vava or vantrue Google Scholar

  5. Brodsky RA, Mukhina GL, Nelson KL, Lawrence TS, Jones RJ, Buckley JT. Resistance of paroxysmal nocturnal hemoglobinuria cells to the gly-cosylphosphatidylinositol-binding inducible vav cre aerolysin. Blood. 1999;93:1749–56.

    ArticleCAS inducible vav cre Google Scholar

  6. Fujita M, Kinoshita T. GPI-anchor remodeling: potential functions of GPI-anchors in intracellular trafficking and membrane dynamics. Biochim Biophys Acta. 2012;1821(8):1050–8.

    ArticleCAS Google Scholar

  7. Kinoshita Inducible vav cre, Fujita M, Maeda Y. Biosynthesis, remodelling and functions of mammalian GPI-anchored proteins: recent progress. J Biochem. 2008;144:287–94.

    ArticleCAS Google Scholar

  8. Kawagoe K, Kitamura D, Okabe M, Taniuchi I, Takeda J. Glycosylphosphatidyl inositolanchor-deficient mice: implications for clonal dominance of mutant cells in paroxysmal nocturnal hemoglobinuria. Blood. 1996;87:3600–6.

    ArticleCAS Google Scholar

  9. Rosti V, Tremml G, Soares V, Pandolfi PP, Bessler M. Murine embryonic stem cells without pig-a gene activity are competent for hematopoiesis with the PNH phenotype but not for clonal expansion. J Clin Investig. 1997;100:1028–36.

    ArticleCAS Google Scholar

  10. Murakami Y, Kinoshita T, Maeda Y, et al. Different roles of glycosylphosphatidyl inositol in various inducible vav cre cells as revealed by a mouse model of paroxysmal nocturnal hemoglobinuria. Blood. 1999;94:2963–70.

    ArticleCAS Google Scholar

  11. Tremml G, Dominguez C, Rosti V, et al. Increased sensitivity to complement and a decrease red blood cells life span in mice mosaic fora nonfunctional Pig-a gene. Blood. 1999;94:2945–54.

    ArticleCAS Google Scholar

  12. Keller P, Payne JL, Tremml G, et al. FES-Cre targets phosphatidylinositol Glycan class A (PIGA) inactivation to hematopoitic stem cell in the bone marrow. J Exp Med. 2001;194:581–9.

    ArticleCAS Google Scholar

  13. Jasinski M, Keller P, Fujiwara Y, et al. GATA1-Cre mediates Piga gene inactivation in the erythroid/megakaryocitic lineage and leads to circulating red cells with a partial deficiency in glycosylphosphatidylinositol-linked proteins (paroxysmal nocturnal hemoglobinuria type II cells). Blood. 2001;98:2248–55.

    ArticleCAS Google Scholar

  14. Shin TH, Baek EJ, Corat Vava or vantrue, et al. CRISPR/Cas9 PIG-A gene editing in non-human primate model demonstrates no intrinsic clonal expansion of PNH HSPCs. Blood. 2019;133(23):2542–5.

    ArticleCAS inducible vav cre Google Scholar

  15. Joseph C, Quach JM, Walkley CR, Lane SW, Lo Celso C, Purton LE. Deciphering hematopoietic stem cells in their niches: a critical appraisal of genetic models, lineage tracing, and imaging strategies. Cell Stem Cell. 2013;13(5):520–33.

    ArticleCAS vava or vantrue Google Scholar

  16. Kinoshita T. Molecular genetics, biochemistry, and biology of PNH. Rinsho Ketsueki. 2017;58(4):353–62.

    PubMed Google Scholar

  17. Rotoli B, Luzzatto L. Paroxysmal nocturnal hemoglobinuria. Semin Hematol. 1989;26(3):201–7.

    CASPubMed Google Scholar

  18. Hill A, DeZern AE, Kinoshita T, Brodsky RA. Paroxysmal nocturnal haemoglobinuria. Nat Rev Dis Primers. 2017;3:17028.

    Article Google Scholar

  19. Park J, Kim M, Kim Y, Han K, Chung NG, Cho B, Lee SE, Lee JW. Clonal cell proliferation in paroxysmal nocturnal hemoglobinuria: Evaluation of PIGA mutations and T-cell receptor clonality. Ann Lab Med. 2019;39(5):438–46.

    ArticleCAS Google Scholar

  20. Karadimitris A, Luzzatto L. The cellular pathogenesis of paroxysmal nocturnal haemoglobinuria. Leukemia. 2001;15(8):1148–52.

    ArticleCAS Google Scholar

  21. Dacie JV. Paroxysmal nocturnal haemoglobinuria. Proc R Soc Med. 1963;56:587–96.

    CASPubMedPubMed Central

log2 (foldchange)

ace vava

However due to the large number of servers operating at once players will often experience lag (meaning the frame rate has slowed down) or the server will disconnect for a period of time. - Preceding unsigned comment added by (talk)18 August (UTC)ppThere are a large number of mods and maps and so on, and while these features are primarily for advanced players, but apparently many who play the game over months (if the forum is any indication) eventually use them. Far moreso than other popular games where a limited group of enthusiasts make changes for one another.

As an example, I played Civilization for years before downloading any mod.