doi:10.1016/j.jmb.2004.06.030 J. Mol. Biol. (2004) 341, 635–649 In Vivo Selection of Combinatorial Libraries and Designed Affinity Maturation of Polydactyl Zinc Finger Transcription Factors for ICAM-1 Provides New Insights into Gene Regulation Laurent Magnenat, Pilar Blancafort and Carlos F. Barbas III* The Skaggs Institute for Chemical Biology and the Department of Molecular Biology, The Scripps Research Institute, 10550 North Torrey Pines Road, La Jolla, CA 92037 USA Zinc finger DNA-binding domains can be combined to create new proteins of desired DNA-binding specificity. By shuffling our repertoire of modified zinc finger domains to create randomly generated polydactyl zinc finger proteins with transcriptional regulatory domains, we developed large combinatorial libraries of zinc finger transcription factors (TFZFs). Millions of TFZFs can then be simultaneously screened in mammalian cells. Here, we successfully isolated specific TFZFs that significantly positively and negatively modulate the transcription of the ICAM-1 gene in primary and cancer cells, which are relevant to ICAM-1 biology and tumor development. We show that TFZFs can work in a general and in a cell-type specific manner depending on the regulatory domain and the zinc finger protein. We show that a TFZF that interacts directly with the ICAM-1 promoter at an overlapping NF-kB binding enhancer can overcome or synergistically cooperate with NF-kB induction of ICAM-1. For this TFZF, rational design was used to optimize the binding of the zinc finger protein to its DNA element and the resulting TFZF demonstrated a direct correlation between increased affinity and efficiency of target gene regulation. Thus, combining library and affinity maturation approaches generated superior TFZFs that may find further applications in therapeutic research and in ICAM-1 biology, and also provided novel mechanistic insights into the biology of transcription factors. Transcription factor libraries provide genome-wide approaches that can be applied towards the development of TFZFs specific for virtually any gene or desired phenotype and may lead to the discovery of new genetic functions and pathways. q 2004 Elsevier Ltd. All rights reserved. *Corresponding author Keywords: polydactyl zinc finger; designer transcription factor; ICAM-1 regulation; in vivo library selection; affinity maturation Introduction The modularity of the zinc finger (ZF) domains Abbreviations used: ZF, zinc finger; TFZFs, ZF transcription factors; ZFPs, ZF proteins; IRES, internal ribosome entry site; GFP, green fluorescent protein; KRAB, Krüppel-associated box; SID, mSin3 interaction domain; HUVEC, human umbilical vein endothelial cells; ChIP, chromatin immunoprecipitation; MFI, mean fluorescence intensities; EMSA, electrophoretic mobilityshift assays; EGF, epidermal growth factor; ICAM-1, intercellular cell adhesion molecule. E-mail address of the corresponding author: carlos@scripps.edu allows for the development of ZF transcription factors (TFZFs) that control the expression of genes of biological and therapeutic interest. Prototypical ZF domains bind 3 bp of DNA sequences through the formation of specific contacts primarily within the major groove of the DNA. By using selective strategies, our laboratory and others have successfully changed the sequence specificity of ZF domains in a directed fashion and have generated polydactyl zinc-finger proteins for targeting unique sites within complex genomes.1 – 5 When fused to transcription activation or repression domains, designed ZF proteins (ZFPs) become regulators of the transcriptional activity of target genes in cultured cells and in living plants and animals.5 – 11 0022-2836/$ - see front matter q 2004 Elsevier Ltd. All rights reserved. 636 Directed artificial gene regulation with rationally designed ZFPs can be limited by a lack of information concerning the target gene, including chromatin structure, the presence of endogenous transcription factors and DNA accessibility. As an alternative to the design and testing of singular ZFPs, we recombined our set of predefined ZF domains to construct random libraries of threeand six-ZF proteins. When attached to the desired effector domain, the large libraries of new polydactyl-ZF DNA binding proteins become genomewide tools that can be screened in vivo (in this context referring to events occurring in living cells but potentially in whole organisms) by selection in mammalian cells for the discovery of novel functional transcription factors. We recently reported preliminary studies concerning the construction and screening of TFZF libraries for the selection of VE-cadherin gene regulators.12 The transcriptional regulation of the gene encoding for the intercellular cell adhesion molecule 1 (ICAM-1, CD54) is dynamic and is implicated in biology and in a variety of diseases. Disorders associated with ICAM-1 deregulation include, malignancies, inflammatory disorders, atherosclerosis, ischemia, neurological disorders and organ transplantation. ICAM-1 is expressed at a basal level in many cell types, including leukocytes and endothelial cells, and binds to the b2 integrins present on the cell surface of leukocytes.13 This interaction promotes adhesion and signaling for transendothelial migration of leukocytes and for T-cell co-activation during inflammatory and immune responses.14 Significantly, ICAM-1 transcription is spatiotemporally regulated in endothelial and cancer cells during tumor angiogenesis, metastasis and progression. Given the diverse roles of ICAM-1 in biology, directed regulation of ICAM-1 expression with novel TFZFs might be an important tool in vivo for the development of anti-inflammatory and anticancer therapies. Here, we have chosen ICAM-1 regulation as a model system and report a new approach to the discovery and optimization of transcriptional regulators. This study led to the development of a set of ICAM-1 regulators (CD54-TFZFs), that are able to significantly up-regulate or completely suppress ICAM-1 expression in primary cells and a variety of cell lines of special interest for ICAM-1 biology. Moreover, we demonstrate that one of the selected CD54-TFZFs interacts directly with the ICAM-1 promoter at a site, normally known to confer responsiveness to natural inducers via NF-kB signaling pathway. The other CD54-TFZFs may regulate via unknown DNA elements, genes and genetic pathways involved in ICAM-1 expression. In order to understand the generality and the particulars of this approach, the activities of the TFZFs were evaluated in different contexts by testing different cell-types, by comparing regulatory domains, and by modifying zinc finger characteristics, including DNA-binding affinity and specificity. Thus, in this study, we detail a Novel ICAM-1 Zinc Finger Transcription Factors powerful strategy for generating new transcription factors for the potent activation and repression of endogenous genes. Additionally, this study brings valuable new mechanistic insight into the biology of TFZFs. These insights will be important for others interested in either engineered or natural transcription factors. Results Selection of TFZFs libraries in mammalian cells for ICAM-1 regulators Whereas three-ZF proteins recognize a 9 bp site with affinities in the nanomolar range, proteins containing six-ZF domains typically bind to 18 bp sequences with better affinities.2,5,15 Therefore, in order to target sites that are in principle unique within the human genome, we developed and used a library of artificial transcription factors containing shuffled 6ZF modules with the canonical TGEKP linker as a connector between the ZF domains wherein the resulting DNA-binding protein is fused to the VP16-derived transactivation domain VP6412 (Figure 1(a)). Most of the domains included in the library were selected in vitro by phage display and optimized by sitedirected mutagenesis for specific binding to each of the possible GNN subsites.3,16 Additional ANN and TNN domains4,17 were also utilized to prepare a six-finger library with a diversity of 8.42 £ 107 proteins, approximately 2000 times as many transcription factors as there are genes in the human genome. The 6ZF-VP64 library was introduced into A431 cells by retroviral infection using the pMX-IRESGFP retroviral vector, which expresses a single bicistronic message for the translation of an effector gene and, from an internal ribosome entry site (IRES), the green fluorescent protein (GFP).7,18 Since both proteins are coded on the same RNA, GFP expression is used as an indicator of infection efficiency and ZF expression. For screening the pMX-6ZF-VP64 library for ICAM-1 regulators, infected cells showing ICAM-1 immunofluorescence up-regulation were isolated using flow cytometry. After cell sorting, the recovered cells were harvested to prepare genomic DNA. The retrovirally integrated ZF pools were PCR amplified and recloned into pMX-VP64-IRES-GFP vector for further rounds of selection (Figure 1(b)). During the three rounds of selection, the ICAM-1 cell-surface expression gradually increased and plateaued. A total of 35 individual CD54-TFZFs clones were then screened. Among the most potent transactivators of ICAM-1, four 6-ZF proteins, CD54-3, CD5413, CD54-30 and CD54-31, were independently cloned multiple times and reproducibly up-regulated endogenous ICAM-1 expression in A431 cells (Figure 1(c)). The expression level of the endogenous ICAM-1 protein, which is already present at moderate levels on normal A431 cells, Novel ICAM-1 Zinc Finger Transcription Factors 637 Figure 1. TFZF library, selection strategy and isolation of TFZF clones. (a) Shuffling of ZF domains generates a large library of 6-ZFPs. Fusion of the library to transcriptional effector domains allows for the screening of TFZFs regulating virtually any gene. (b) ICAM-1 up-regulators were isolated by FACS sorting of GFPþ and ICAM-1þ cell population from the gated region R1 and by recovery from integrated retroviral 6ZF DNA pools and subsequent rounds of selection. The relative increase in ICAM-1 mean fluorescence intensity (MFI) during three rounds of selection is compared to mock infected cells (pcDNA) or cells infected with the unselected library (6ZFlib). (c) ICAM-1 cytometric profiles of cells infected with the cloned retroviral CD54-transactivators (bold line) were compared to cells infected with the unselected pMX-6ZF-VP64 library (thin line). The basal autofluorescence of A431 cells is represented by a broken line. was increased two to four times as compared to cells infected with the unselected TFZFs library (Table 1). Interestingly, up-regulation was not correlated with the level of GFP expression (data not shown), suggesting that CD54-TFZFs clones may have different intrinsic efficiencies as regulators. The polydactyl ZF domains were sequenced and their predicted target sites were deduced from our database of ZF domains and corresponding 3 bp subsites. The CD54-6ZF proteins, expressed and purified as maltose-binding protein (MBP-CD54) fusions, bound their predicted 18 bp target sequences with affinities ranging from 1 nM to 10 nM as determined in vitro by EMSA (Table 1). Effector domain swapping in CD54-TFZFs for ICAM-1 repression In order to direct repression of ICAM-1, CD546ZF proteins were subcloned into the pMX retroviral vector as fusions to the transcriptional repressors Krüppel-associated box (KRAB) and the mSin3 interaction domain (SID).6,19,20 The resulting pMX-CD54-KRAB and pMX-CD54-SID constructs were tested for repression of endogenous ICAM-1 expression in A431 carcinoma, C8161 melanoma and Kaposi’s sarcoma SLK cells (Figure 2). The KRAB domain only functioned as an ICAM-1 repressor in fusions with ZFPs CD54-13 and CD54-31. Potent ICAM-1 knock-down was achieved with CD54-31-KRAB in three cell lines (9 – 1% ICAM-1 remaining), whereas 50 – 13% ICAM-1 still remained with the same ZFP in combination with the SID domain. On the other hand, the SID domain worked as repressor with all four CD54-6ZF at variable degrees depending on the cell line. For instance, CD54-30 was able to significantly repress ICAM-1 down to 15% of the native level, but only with SID and only in C8161 cells. Also, CD54-3 worked exclusively with SID. However, CD54-13 paired equally with both repressors. The CD54-TFZFs transcriptional levels measured by semi-quantitative RT-PCR varied slightly depending on the nature of zinc finger and 638 DCRDLAR RSDDLVR QAGHLAS CD54-31Opt The fold activation of the endogenous ICAM-1 gene in A431 cells and dissociation constants are shown. a Zinc finger helices are positioned in the anti-parallel orientation (C-terminal-F6 to F1-N-terminal) relatively to the DNA target sequence. Amino acid position 21 to þ 6 of each DNA recognition helix is shown. b Predicted target DNA sequences are presented in the 50 to 30 orientation. c Fold increase ICAM-1 expression in A431 cells. Mean fluorescence intensity (MFI) corrected for background auto-fluorescence and standardized to the ICAM-1 intensity of A431 cells infected with the unselected 6ZF-VP64 library ( ¼ 1). Average of relative MFI from 2 to 11 independent experiments (statistical significance of observed differences was determined using the Student’s t-test, p , 0:06). d Dissociation constant ðKd Þ determined by gel shift assay. Data represent the average of two to four independent EMSA experiments. Zinc finger domains used for CD54-31 optimization (CD5431Opt) are presented in bold. Nucleotide positions that do not match the predicted target sequence are in lowercase letters. 9.1 £ 7.1 5.1 8.5 1 3.1 0.16 3.6 £ 2.2 £ 4.0 £ 3.6 £ GCC-30 GTA-30 GCA-30 GCC-30 GCC-30 GCC-30 GTT GAC GCC GGG GTA GGA AAA GCG gAA GCG GAA GCG – – – – – – 50 -AAA GTT AAA 50 -GAC GGT AAA 50 -GAA GTT GTA 50 -TGA TGA GTT 50 -Tcc gGA GcT 50 -Tcc GGA GCT DCRDLAR QSSSLVR QSGDLRR DCRDLAR QRANLRA DPGNLVR QSSNLVR QAGHLAS CD54-3 CD54-13 CD54-30 CD54-31 TSGSLVR QRANLRA TSGSLVR TSGHLVR QRANLRA DCRDLAR TSGSLVR QSSSLVR QSSSLVR QAGHLAS TSGSLVR QRANLRA Natural ICAM-1 promoter site (pro-220): QRAHLER TSGELVR QSSNLVR DPGNLVR RSDKLVR QRAHLER RSDDLVR Half-site 1 N-term F1 F2 F3 F4 F5 C-term F6 6ZF Zinc finger helicesa Table 1. DNA interacting helices of 6ZF clones activating ICAM-1 and their predicted 18 bp target sites Target sitesb Half-site 2 Fold (MFI)c Kd (nM)d Novel ICAM-1 Zinc Finger Transcription Factors repression domains and correlated well with the fluorescence levels of the bicistronically co-expressed GFP marker in A431 cells (Figure 2(a)). However, the variations of both TFZFs mRNA and GFP marker did not correlate with the efficiency of ICAM-1 repression by the different CD54-TFZFs and could not account for the striking differences observed above (Figure 2(b) –(d)). Thus, the choice of repressor domain is important depending on the ZF protein and the cell line to repress ICAM-1. We chose CD54-31-KRAB as a general repressor for further studies. Artificial TFZFs function efficiently in primary cells Artificial TFZFs perform well in cancer and transformed cells, but their activity in primary human cells remains to be fully explored. Therefore, we assessed CD54-31 TFZFs in primary cells and telomerase (hTERT)-immortalized primary cells, shown to maintain a normal karyotype and phenotype.21,22 The effect of CD54-31 fusion proteins was considerable in human umbilical vein endothelial cells (HUVEC), where the ICAM-1 level was reduced by 83% in combination with KRAB and increased 63-fold with VP64 (Figure 3). In immortalized human primary mammary epithelial cell line (hTERT-HME1) and in human hTERT-fibroblasts, where ICAM-1 was present at much higher extent, regulation was also effective (Figure 3). These results demonstrate the wide potential of the TFZFs in normal cells. ICAM-1 promoter scanning and the search for the CD54-31-binding site TFZFs act by binding to DNA and recruiting different factors of the transcription machinery. In order to determine if the selected CD54-TFZFs bound directly to the ICAM-1 promoter or may have acted indirectly through the activation of other genes involved in the expression of ICAM-1, CD54-6ZF coding DNAs were subcloned into the pcDNA-VP64 transient expression vector for activation studies with a luciferase reporter construct driven by a 1.6 kb fragment of the ICAM-1 promoter (pGL3-ICAM-1). Two of the CD54-TFZFs, CD54-3 and CD54-31, were able to significantly up-regulate luciferase activity in transient transfection assays in A431 cells (Figure 4(a)). Comparison of the promoter sequences and the predicted 18 bp target sequence of CD54-31 ZFP suggested several potential target sites. In order to determine the site the TFZF functions through, purified MBP-CD54-31 ZF fusion protein was tested in an ELISA assay for binding to the candidate target DNAs. Whereas the CD54-31 protein bound preferentially to its predicted DNA sequence (cd54-31) at concentrations in the nanomolar range, it also bound effectively to an 18 bp promoter site, ICAM-1 pro-220 (Figure 4(b)). This 18 bp sequence shared 13 nt with the predicted sequence (Table 1) and was found 639 Novel ICAM-1 Zinc Finger Transcription Factors Figure 3. ICAM-1 regulation in primary cells mediated by CD54-TFZFs. HUVEC cells, telomerase immortalized primary mammary epithelial cells hTERT-hME1 and hTERT-fibroblast retrovirally expressing CD54-31 ZFP fused to either VP64 (red line) or KRAB (blue line) domains were compared to the normal ICAM-1 expression of cells infected with the unselected pMX-6ZF library (green line) and the basal autofluorescence (broken line). 220 nt upstream of the initiation of translation of ICAM-1.23 In agreement with this result, the affinity of the MBP-CD54-31 protein for the predicted cd54-31 oligonucleotide was 1 nM and 3.1 nM for the ICAM-1 pro-220 DNA (Table 1). Significantly, this promoter sequence overlaps an enhancer element, shown to bind to NF-kB and convey responsiveness to tumor necrosis factor-a (TNFa), interleukin 1b (IL-1b) and other factors.13 Natural target site validation and CD5431 optimization When the ICAM-1 pro-220 site was replaced by the target site of an unrelated 6-ZFP E2C in the luciferase construct (pGL3-E2C/ICAM-1), CD5431-VP64 was no longer able to efficiently up-regulate luciferase activity. However, the mutant promoter was nonetheless inducible by the E2C-VP64 control protein, showing that the substitute e2c sequence was not detrimental to the reporter construct, further indicating that ICAM-1 pro-220 is the target site of CD54-31 (Figure 4(c)). Because the 18 bp target site found in the ICAM-1 promoter was not optimal for the CD54-31 protein, the 6ZF protein was optimized for binding the natural promoter site using rational design and protein grafting of specific DNA-binding a-helices6 (Table 1). For instance, the ZF domains selected in Figure 2. ICAM-1 repression and relative CD54-TFZFs expression in different cancer cell lines. (a) Correlation of mean fluorescence intensities (MFI) of the coexpressed GFP marker (black bars, average of three independent experiments) with TFZF – KRAB mRNA levels measured by RT-PCR (white bars) in A431 cells. Values are represented relatively to the CD54-3-KRAB levels ( ¼ 1). Decrease in ICAM-1 MFI in (b) A431 carcinoma, (c) C8161 melanoma and (d) Kaposi sarcoma SLK cells infected with CD54-3, CD54-13, CD54-30 and CD54-31 6-ZFPs fused to either KRAB (black bars) or SID (white bars) repression domains. Mean fluorescence intensities (MFI) are represented as percent ICAM-1 remaining and standardized to normal ICAM-1 levels in cells infected with the unselected pMX-6ZF library (6ZFlib ¼ 100% ICAM-1 remaining) and to basal autofluorescence ( ¼ 0% ICAM-1) as determined by FACS. In parenthesis are indicated the relative bicistronically expressed GFP levels (MFI) corrected for background autofluorescence for each CD54-TFZFs. 640 Novel ICAM-1 Zinc Finger Transcription Factors Figure 4. ICAM-1 promoter regulation, natural target site determination and CD54-31 optimization. (a) ICAM-1 promoter activation in transient luciferase reporter assays. Normalized luciferase activity of pGL3-ICAM-1 reporter induced by CD54-transactivators was compared relatively to pcDNA vector control ( ¼ 1). (b) Promoter scanning ELISA. The DNA-binding intensity of purified MBP-CD54-31 protein for potential natural ICAM-1 promoter sites is presented relatively to the predicted target site cd54-31 ( ¼ 1). (c) Mutant reporter assay. The pGL3-E2C/ICAM-1 mutant reporter construct was tested for lack of transactivation with the CD54-31-VP64 construct. The E2C-VP64 positive control was known to activate the e2c target site in the ErbB2 gene.6 (d) Effect of CD54-31 optimization on the transactivation of the ICAM-1 promoter. Normalized luciferase activity of pGL3-ICAM-1 reporter induced by CD5431 and CD54-31Opt transactivators is presented as described in (a). (e) DNase I footprints of purified CD54-31 (left gel) and CD54-31Opt (right gel) MBP fusion proteins on the ICAM-1 promoter. DNA incubated with tenfold dilutions of ZFP (1000 –1 nM) was run in parallel to DNA with DNase I only (DNase I), DNA alone (Control), and chemical sequencing of the 200 bp DNA probe (G þ A ladder). The ICAM-1 pro-220 target site (capital letters), site of DNase I protection (dotted line) and the overlapping NF-kB enhancer (broken line) are presented. CD54-31 for finger positions F3-AAA (QRANLRA), F4-GTT (TSGSLVR) and F5-TGA (QAGHLAS) were, respectively, shown to bind in vitro to the triplets GNN, GCT and GGA with reduced affinity.3,16 Suitably, those triplets matched the ICAM-1 pro-220 sequence (50 -TCC GGA GCT GAA GCG GCC-30 ). Therefore, three out of six ZF domains were replaced in a newly designed and optimized Novel ICAM-1 Zinc Finger Transcription Factors 641 Figure 5. In vivo binding of CD54-TFZFs to the ICAM-1 promoter and effects on the NF-kB signaling pathway inducing ICAM-1 in response to proinflammatory cytokines. (a) ChIP assay with normal (A431) and TFZFs expressing (31-KRAB and 31Opt-KRAB) cells. Formaldehyde cross-linked chromatin was immunoprecipitated with a TFZF-specific antibody (ZF) or without antibody (No) and analyzed by PCR using primers specific to the ICAM-1 promoter region surrounding the NF-kB enhancer. Total input chromatin (In) was used as positive control. (b)– (e) Inhibition and synergistic up-regulation of NF-kB mediated cytokine induction of ICAM-1 with CD54-TFZFs. FACS analysis of ICAM-1 in HUVEC (b) and (c) and A431 cells (d) and (e) retrovirally infected with CD54-31-KRAB (red line) and CD54-31Opt-KRAB (orange line) TFZFs and subsequently treated with TNF-a (b) or Il-1b (c) – (e). Normal ICAM-1 expression in control cells without (green broken line) and with cytokine treatment (green plain line), and basal autofluorescence (black broken line) are presented. (e) ICAM-1 mean fluorescence intensities of untreated (black bars) and IL-1b (white bars) treated A431 cells infected with CD54-31 and CD54-31Opt zinc finger proteins alone (31- and 31Opt-) or either coupled to KRAB (31-KRAB and 31Opt-KRAB) or VP64 (31-VP64 and 31Opt-VP64) domains. protein (CD54-31Opt) with domains determined to have better specificity in vitro, namely F3-GAA (QSSNLVR), F4-GCT (TSGELVR) and F5-GGA (QRAHLER) (presented in bold in Table 1). Since the finger positions F1-GCC and F2-GCG were optimal and because no domains were available for a F6-TCC subsite, the originally selected DNA binding a-helices were retained at these positions in the protein. The affinity of purified MBP-CD5431Opt fusion protein for the ICAM-1 pro-220 site was determined to be 0.16 nM, approximately 20 times better than the dissociation constant of CD54-31 for the same DNA (Table 1). As a result, CD54-31Opt-VP64 induced twice the level of luciferase activity as the original CD54-31-VP64 construct in transiently transfected A431 cells (Figure 4(d)). Finally, DNase I footprinting demonstrated actual binding of both proteins, CD54-31 and CD54-31Opt, to the same ICAM-1 pro-220 promoter sequence in vitro (Figure 4(e)). Together, these results convincingly demonstrate that ICAM-1 pro-220 is the endogenous target site of CD54-31 and that it can be further induced by designed affinity maturation of CD5431Opt. In vivo binding, competition and synergistic cooperation of CD54-TFZFs with endogenous factors at the ICAM-1 promoter A chromatin immunoprecipitation (ChIP) experiment was performed to establish the interaction of the selected and optimized CD54-TFZFs with the ICAM-1 promoter site in vivo (Figure 5(a)). A chromatin fragment comprising the ICAM-1 pro-220 site was efficiently cross-linked by the CD54-31-KRAB and -31Opt-KRAB proteins expressed in A431 cells and co-immunoprecipitated with a polyclonal antibody raised specifically against the framework of our designer zinc finger domains. This framework consist of a consensus peptide sequence derived from 131 zinc finger domains of different origins,24 that was shown to provide better stability and affinity properties to Sp1C, a hybrid zinc finger protein in which only the residues from the three DNA recognition helices of the natural Sp1 transcription factor were grafted into the consensus framework.25 While the same antibody specifically reacted with retrovirally expressed CD54-TFZFs in crude cell extracts on Western blot (data not shown), no 642 Novel ICAM-1 Zinc Finger Transcription Factors Figure 6. Specificity of the CD54-TFZFs. (a) Multitarget specificity in vitro assay. Purified MBP CD54-6ZF fusion proteins CD54-3, CD54-13, CD-54-30 and CD54-31 were tested in a DNA binding ELISA against each corresponding target DNA oligonucleotides, cd54-3, cd54-13, cd54-30 and cd54-31. The CD54-31Opt protein and the ICAM-1 pro-220 promoter target site were also included. The DNA-binding intensity (A405 nm) is presented for each protein relatively to the respective target site ( ¼ 1) as determined from duplicate experiments, for which dilution series of 60 nM protein were used. (b)– (h) Multitarget specificity of CD54-TFZFs in cancer and primary cells. Retrovirally expressed CD54TFZFs, CD54-31Opt (orange line) and CD54-31 (red line) VP64 activators in (b) A431 and (c) HUVEC cells, (d) KRAB repressors in A431 cells, and (f) CD54-3-VP64, (g) CD54-13-VP64 and (h) CD54-30-VP64 (all red line) in A431 cells, were compared by FACS for the regulation of endogenous ICAM-1 and evaluated for regulation of other cell surface markers Epidermal growth factor (EGF), 3-FAL selectin ligand (CD15, FUT4), integrin-a6 (CD49f, ITGA6), leukocyte function-associated antigen (CD58, LFA-3), Apo1-FAS antigen (CD95, TNFRSF6), integrin-b4 (CD104, ITGB4) and vascular endothelial VE-cadherin (CD144, CDH5). Control cells are labeled as in Figure 3. (e) Semi-quantitative RT-PCR analysis of A431 cells retrovirally infected with CD54-TFZF KRAB repressors for specificity of ICAM-1 and ITGA6 regulation at the transcriptional level in relation to TFZFs and GAPDH mRNAs levels. Controls experiments include RNAs from mock infected A431 cells (M), cells infected with the unselected pMX-6ZF library (6ZFlib) and from CD54-31OptKRAB cells in the absence of reverse transcriptase (2). 643 Novel ICAM-1 Zinc Finger Transcription Factors immunoprecipitated DNA fragment was detected by PCR in normal A431 cells, nor without the addition of the antibody. This demonstrates that the immunoprecipitation complex was only formed with our artificial TFZFs but not with endogenous zinc finger containing factors. While CD54-31-KRAB was able to efficiently repress the constitutive expression of ICAM-1 (Figure 2) and bound to a sequence that was overlapping with the binding element of NF-kB, it was interesting to determine the effect on the induced over-expression of ICAM-1 in response to cytokines via NF-kB signaling (Figure 5(b) – (e)). Mutation of this site completely abolished ICAM-1 promoter activation by TNF-a and IL-1b,26 the major inducers of ICAM-1 expression in most cell types which act through NF-kB signaling pathway. Amongst all the ICAM-1 inducing factors assayed, TNF-a was a better inducer than Il-1b by increasing expression more than 20-fold over the weak constitutive ICAM-1 expression observed in HUVEC cells (Figure 5(b) and (c)) and IL-1b was the most potent inducer in A431 cells (Figure 5(d) and (e)). CD54-31 and -31Opt TFZFs coupled to VP64 activators up-regulated ICAM-1 through a range far exceeding the induction provided by natural factors (Figure 5(e)). Most importantly, when coupled to KRAB repressors they were still able to completely repress the inducible and constitutive ICAM-1 expression in the presence of cytokines, indicating their ability to either compete or overcome NF-kB regulation in both cell lines. Similarly, total inhibition of ICAM-1 induction was also seen in A431 and HUVEC with other factors such as lipopolysaccharide, phorbol 12-myristate-13-acetate, and interferon gamma, which signals through other transcription factors that bind to ICAM-1 regulatory elements different from the NF-kB enhancer (data not shown). Interestingly, the same zinc finger proteins coupled to the VP64 activation domain, in combination with Il-1b together increased ICAM-1 expression greater than the additive increase provided by the factors individually. Zinc finger proteins lacking the effector domain had no measurable effect (Figure 5(e)). Specificity of CD54-TFZFs and activity in biologically relevant cell lines As the specificity and modularity of individual ZF domains assembled in a polydactyl ZF protein can vary, it was important to evaluate the specificity of the CD54-6ZF proteins. Initially, purified fusion proteins were assessed in a multi-target specificity ELISA for binding to each of the predicted or natural target sites. As shown in Figure 6(a), each protein bound specifically to its corresponding DNA hairpin oligonucleotide. Note that the CD5431Opt protein was redirected towards binding to the ICAM-1 pro-220 target DNA with a fivefold preference over the originally predicted cd54-31 target site. Next, the retrovirally expressed CD54-VP64 TFZFs were tested for their ability to alter ICAM-1 expression. Study of cytometric profiles of ICAM-1 and seven other cell surface markers in A431 and HUVEC cells infected with CD54-31-VP64 and CD54-31Opt-VP64 revealed that both TFZFs reproducibly and preferentially up-regulated ICAM-1 compared with the other markers (Figure 6(b) and (c)). Further, CD54-31Opt was more than twice as potent as CD54-31 in both cell lines (Table 2), thus reaching one order of magnitude increase in ICAM-1 fluorescence signal in A431 and two in HUVEC. Optimization of CD54-31 also corrected the non-targeted regulation found with CD54-31 for one marker (CD144) in A431 and another in HUVEC (CD95), while slightly affecting others (CD104 and CD58). The specificity of the KRAB effector variants was also tested in A431 cells (Figure 6(d)), where both selected and optimized CD54-31 TFZFs repressed completely ICAM-1 but did not affect the other genes. A semi-quantitative RT-PCR was performed (Figure 6(e)) and only the ICAM-1 mRNA was Table 2. Fold activation and repression ICAM-1 expression with CD54-31 and CD54-31Opt TFZFs in cell lines CD54-31 Type Cancer Primary Cell line Colon Colon Colon Breast Breast Breast Melanoma Epidermoid Endothelium Lim1215 HT29 SW1222 T47D SKBR3 MD-AMB-435S C8161 A431 HUVEC a b Relative ICAM-1 fluorescence VP64 85 62 1 33 4 627 31 93 6 1.9 7.3 137.0 13.0 59.0 (14.5) 3.1 22.1 (0.27) 3.6 (1) 63.0 (2.83) CD54-31Opt c b KRAB VP64 (Fold opt.d) KRABc 3 (2.8) 5 (1.4) 23 (8) 3.4 10.1 142.0 9.6 58.0 4.5 59.0 9.1 135.0 1.8 £ 1.4 £ 1.0 £ 0.7 £ 1.5 £ 1.0 £ 2.7 £ 2.5 £ 2.1 £ 10 6 13 a Normal ICAM-1 mean fluorescence intensity is corrected for background and standardized to the autofluorescence ( ¼ 1). Average value from two to seven experiments. b Fold increase in ICAM-1 expression. Mean fluorescence intensity with VP64 transactivators relatively to cells infected with unselected 6ZF-VP64-library ( ¼ 1). In parentheses is presented the standard deviation of two independent experiments. c Percentage remaining ICAM-1 mean fluorescence intensity with the KRAB repressors relatively to cells infected with unselected 6ZF KRAB-library ( ¼ 100%) and autofluorescence ( ¼ 0%). In parentheses is presented the standard deviation of two independent experiments. d Additional fold VP64 activation obtained with CD54-31Opt as compared to CD54-31. 644 Novel ICAM-1 Zinc Finger Transcription Factors Figure 7. ICAM-1 regulation in different cell types. Activators CD54-31-VP64 (red line) and CD54-31Opt-VP64 (orange line) and repressors CD54-31-KRAB (light blue line) and CD54-31Opt-KRAB (dark blue line) were retrovirally expressed in cells derived from colon cancer (Lim1215, HT29 and SW1222), breast cancer (T47D, SKBR3 and MD-AMB435S), melanoma (C8161), carcinoma (A431) and primary human umbilical vein endothelial cells (HUVEC). Control cells are labeled as in Figure 3. repressed with the CD54-31 and 31-Opt KRAB TFZFs, whereas ITGA6 and GAPDH retained normal levels of mRNA. Of the remaining CD54-TFZF regulators, CD54-3 presented the best overall specificity towards ICAM-1, while CD54-13 and CD54-30 presented varying degrees of specificity (Figure 6(f)–(h)). Overall, CD54-TFZFs that showed regulation of the ICAM1 promoter in transient assays (namely, CD54-3, -31 and -31Opt) were specific to ICAM-1 and showed no or little change in expression for the other markers. Since ICAM-1 transcriptional regulation plays major roles in biology and cancer development in endothelial and cancers such as breast, colon and melanoma, and the major transcription-regulatory response element is the kB enhancer that overlaps with the ICAM-1 pro-220 target site, the ability to effectively interfere with the ICAM-1 expression in this context is of considerable interest. Therefore, the activity of CD54-31 and its optimized version was examined for positive and negative regulation of ICAM-1 in relevant cell lines (Figure 7 and Table 2). Both regulators acted equally as repressors with the KRAB domain in A431 carcinoma and in C8161 melanoma and HUVEC cells, where ICAM-1 is moderately expressed as determined by FACS. Levels of mean fluorescence intensity were reproducibly knocked-down over 90%. When used as VP64-based transactivators, CD54-31Opt was twofold more potent in ICAM-1 activation in these cell lines, resulting in 135 and 59 times more ICAM-1 fluorescence in HUVEC and melanoma, respectively. In breast and colon cancer, ICAM-1 may function as a suppressor of tumor progression by promoting necessary interactions with the immune system. ICAM-1 levels in related cell lines were increased by the TFZFs several fold in colon (Lim1215 and HT29) and breast (MD-AMB-435S and T47D) cell lines already showing significant expression, and were increased by up to two orders of magnitude in breast (SKBR3) and colon (SW1222) cell lines, lines that normally show little ICAM-1 expression by FACS. Discussion In the light of the results presented here, we achieved successful selection of specific regulators of ICAM-1 from a library of nearly 108 6ZF transcription factors. First, this strategy overcomes the problems of chromatin structure and accessibility of the target gene and has a great potential for discovery of genes and genetic pathways. Second, considering TFZF biology itself, we found that the Novel ICAM-1 Zinc Finger Transcription Factors selected C54-TFZFs up- and down-regulated ICAM1 in a broad range of cells, including established cancer cell lines, and that TFZFs can also function in primary cells, which is of particular relevance for the biological activity of TFZFs in living organisms. Accordingly, TFZFs can work in a general manner, as seen in the CD54-31 protein, but also in a more complicated fashion. For instance, by doing comparative studies of the SID and KRAB repression domains, we demonstrate for the first time the generation of cell-type specific TFZFs and TFZFs specific to a regulatory domain. CD54-31KRAB most consistently provided full repression, whereas CD54-30-SID efficiently repressed ICAM-1 in C8161 melanoma cells only, and CD543 paired exclusively with the SID domain in the three cell lines tested. Modest differences in the level of expression of the TFZFs seemed not to be indicative of effectiveness, e.g. the best expressing CD54-3-KRAB TFZFs in the three cell lines was inactive, whereas CD54-3-SID which was expressed at , 50% the level of the KRAB factor repressed well. This may be due to the fact that TFZFs are typically expressed in excess compared to their typical targets which are found in two copies per nuclei. Consequently, engineered TFZFs that activate the target endogenous gene, may not necessarily work as a repressor, i.e. there is more at work than simple chromatin accessibility, a requirement that has been heretofore stressed in the literature as perhaps the only requirement for regulation. Accordingly, the distribution of regions accessible to DNase I in the promoter of the VEGF-A locus appeared to be specific to the cell type and was essential in the design of TFZFs targeting VEGF-A.9,27 The differences in activity may also reflect variations in the availability of interacting accessory factors among cell types and their localization on the chromosomes. Accordingly, complex regulation mechanisms involving abundant signaling pathways and transcription factors in ICAM-1 expression have been revealed to be cell-type specific.28 Thus, the unpredictable combination of parameters including zinc finger, DNAbinding element, regulatory domain and cell type, show us a mechanism of action of the TFZFs that is more complex than anticipated. These results better reflect the in vivo situation where chromatin structure is spatio-temporally dynamic and there is a differential expression of factors of the transcriptional machinery. These results together with our further study of ICAM-1 promoter regulation suggest that targeting the proximal promoter of the target gene allows for general activation and repression, whereas targeting sites further away or through indirect regulation of intermediate genes could provide more unique types of regulation. Two of the CD54-TFZFs, CD54-3 and -31, interacted directly with the transcription of the 1.6 kb ICAM-1 promoter and showed better gene specificity than the others studied. This correlation leads us to speculate that CD-54-13 and CD54-30, which did not interact with the reporter construct 645 and appeared to be less gene-specific, may bind less specifically and farther away from the ICAM-1 promoter, or may act by indirect means, e.g. by regulating other genes involved in ICAM-1 expression. It is conceivable that targeting an upstream regulator of ICAM-1 may also result in changes of other genes involved in the same pathway or function, such as the adhesion molecules tested in the gene specificity experiment. Also, selected CD54-TFZFs could regulate genes involved in described ICAM-1 post-transcriptional regulation mechanisms controlling RNA stability, cellsurface localization or proteolytic cleavage from the cell surface,13 or other protein internalization and specific proteolysis mechanisms. CD54-13 could be such a candidate since it represses ICAM-1 cell surface expression consequently (Figure 2), does not seem to affect ICAM-1 mRNA and is not the most gene-specific CD54-TFZF (Figure 6). Blast searches did not reveal near perfect matches in either 100 kb of the ICAM-1 loci, or in the human genome. Previous studies have shown that 6ZFPs provide for higher affinity and specificity compared to 3ZFPs,29 and provided endogenous regulation of the ErbB2 gene, but not the ErbB3 gene, which shared 15 out of 18 identical base-pairs in their respective target sites.7 The selected ZFs bound their predicted target sequences with better than 10 nM affinity. Quite unexpectedly, we found that a 6ZF protein like CD54-31 can regulate a target site through imperfect recognition of 18 bp target sites if they are bound with sufficient affinity, a situation that may reflect the mechanism of general natural transcription factors. However, optimization of the TFZF DNA binding domains using rational design provided a 20-fold increase in affinity to its DNA element. This translated into substantial increases in transient reporter activation and endogenous ICAM-1 regulation. This is the first direct evidence that affinity modulation of a TFZF can be used to modulate the extent of endogenous gene regulation once threshold-binding affinity has been obtained. Further, increased affinity is accompanied by increased specificity. We can speculate that, despite the uniqueness factor of the target site, this may be a mechanism that differentiates general transcription factors from specific ones. Finally, regarding the ICAM-1 target gene biology, the target site cd54-31 was identified on a NF-kB binding element of the ICAM-1 promoter and was verified by promoter site deletion and DNase 1 footprinting in vitro and chromatin immunoprecipitation in vivo. Whereas the nearby Sp1 site is responsible for constitutive ICAM-1 promoter activity, the overlapping NF-kB enhancer confers responsiveness to TNF-a, IL-1b and other signaling molecules.30 Transcription factor complexes of NF-kB, RelA and c-Rel dimers were shown to bind to this element.13 The CD54 regulators described here likely impact the effect of natural factors. Our results indicate that the CD54-31 and CD54-31Opt TFZFs do not compete for the NF-kB 646 enhancer but, when coupled to the N-terminal KRAB repression domain, rather overcome the ability of NF-kB, and other transcription factors, to up-regulate ICAM-1 in response to a variety of inducers, including IL-1b and TNF-a. On the other hand, when coupled to the C-terminal VP64 activation domain, they cooperate synergistically with NF-kB. Note that steric hindrance cannot be the reason for ICAM-1 repression while the TFZF sits 30 of the NF-kB site in an antiparallel orientation relatively to the DNA strand it binds and the C-terminal VP64 activation domain does not seem to interfere with NF-kB dimer binding. Thus, we present evidence that part of the NF-kB enhancer element in the ICAM-1 promoter can also work as repressor element given a specific repressor binds to it and that artificial transcription factors can overcome or synergistically cooperate with endogenous factors. Interestingly, 6-ZFP alone (lacking effector domains) were not able to block either constitutive or induced ICAM-1 expression, despite their subnanomolar binding constants. We also provide a study of the directed gene regulation in a wide panel of cell types, including primary cells and cancer cells that are relevant to its function in the development of tumors. Both CD54-31 and its optimized version, CD54-31Opt, were able to up-regulate endogenous levels of ICAM-1 on the cell surface of colon (up to 142-fold in colon SW1222 cells) and breast (up to 59fold in breast SKBR3 cells) cancer cell lines. During cancer progression, low ICAM-1 correlates with a reduced disease-free state and poor prognosis in colorectal carcinoma patients.31 Accordingly, expression of ICAM-1 in invasive breast cancer reflects low growth potential and good prognosis.32 Reciprocally, ICAM-1 expression in melanoma cells and primary vascular endothelial cells could be completely suppressed and significantly overexpressed (59-fold and 135-fold, respectively). This is of particular importance, since ICAM-1 is specifically down-regulated in tumor endothelial cells during tumor angio-genesis33 and de novo expression of ICAM-1 in melanoma correlates with increased risk of metastasis.34 Therefore, our library and affinity maturation approach provide powerful TFZFs to study ICAM-1 regulation and function by imposed regulation. The main advantage of our strategy is the large complexity of available proteins that can be tested simultaneously, taking into consideration the chromatin accessibility of the targeted endogenous gene and the availability of the transcription machinery in living cells, therefore, overcoming the need for an initial mapping of chromatin-free zones in a target gene of interest.9 Alternatively, our approach could be used to isolate TFZFs associated with specific complex phenotypes beyond the surface expression phenotype selected here. A limitation of this approach is the difficulty in assigning genomic binding sites due to binding of related DNA sequences; however, this has yet to pose significant obstacles. It should be possible to favorably exploit degenerate binding through the Novel ICAM-1 Zinc Finger Transcription Factors use of less specific DNA binding domains. This would allow for more extensive coverage of the genome assuming the corresponding DNA-binding activity was sufficient to impose transcriptional control. This should be feasible, since natural transcription factors bind families of DNA sequences in much the same way. Such approaches might further facilitate the discovery of new genes and pathways involved in the expression of the target gene or phenotype of interest. For instance, the CD54-13 and CD-30 TFZF selected in this study may lead to the discovery of indirect target sites that regulate new genes or genetic pathways involved in the expression of ICAM-1 at the transcriptional and post-transcriptional levels. TFZFs can also be seen as “compact” cDNA libraries when they are expressed with activation domains or their RNAi-like complement when expressed with repression domains. Thus, a wide variety of applications can be envisioned for the transcription factor libraries described here. Materials and Methods Human cell lines A431 epidermoid carcinoma cells and breast cancer cell lines T47D, SKBR3 and MDA-MB-435s were obtained from the American Type Culture Collection (Manassas, VA). Colon cancer cell lines Lim1215, SW1222, HT29 were obtained from the cell bank of the Ludwig Institute for Cancer Research (New York). Kaposi’s sarcoma cell line SLK was provided by R. Pasqualini (University of Texas M.D. Anderson Cancer Center, Houston) with permission from S. LevintonKriss (Tel-Aviv). Melanoma C8161 cells were a generous gift from R. A. Reisfeld (The Scripps Research Institute, La Jolla). Human umbilical vein endothelial cells (HUVEC) were purchased from BioWhittaker (Walkersville, MD) and maintained as indicated by the manufacturer. Immortalized primary mammary epithelial cell line hTERT-HME1 was purchased from CLONTECH (Palo Alto, CA) and hTERT-fibroblasts (J.W. Shay, The University of Texas Southwestern Medical Center, Dallas, TX) were a kind gift of Steven I. Reed (The Scripps Research Institute, La Jolla). The 293-GagPol packaging cell line was obtained from I. Verma (Salk Institute, San Diego). Cells were cultured in RPMI 1640 medium (SLK, SKBR3, Lim1215, SW1222, HT29) or Dulbecco’s modified Eagle’s medium (DMEM) (A431, T47D, MDA-MB-435s, C8161, hTERT fibroblast and 293-gagpol) supplemented with 10% (w/v) FCS and antibiotics. Construction of retroviral and transient expression TFZFs plasmids The pMX-6ZF-VP64 library of 8.4 £ 107 members was constructed as described.12 Retroviral and transient expression of TFZFs, constructs were prepared by subcloning 6ZF-coding DNA into pcDNA- and pMX-effector vectors as described.6,7 For the construction of the CD54-31 and -31Opt proteins lacking an effector domain fusion, the C-terminal AscI-PacI VP64 domain of the pMX-CD54-31 and -31Opt VP64 vectors was removed by replacing the AscI site for a second PacI site using 647 Novel ICAM-1 Zinc Finger Transcription Factors site-directed mutagenesis, followed by PacI digestion and self-religation. The resulting construct conserves the C-terminal hemagglutinin peptide tag in frame with the 6-ZFP for detection of the full-length protein by Western blot (data not shown). Retroviral gene delivery and FACS Screening of the pMX-6ZF-VP64 library for ICAM-1 regulators was performed as described for the isolation of TFZFs specific to the VE-cadherin gene12 with modifications. For optimal infection efficiency, we used a 293-GagPol packaging cell line lacking envelope coding retroviral genes and VSV envelope pseudo-typing by co-transfection of the plasmid pMD.G (obtained from I. Verma, Salk Institute, San Diego). Retroviral infections were set up as following: 3.5 £ 106 293-GagPol cells were plated on a 10 cm polylysine-coated dish and cotransfected with 1.25 mg of pMD.G plasmid encoding the vesicular stomatitis virus-G envelope protein,35 thus conferring a broader host-cell range, a higher viral stability and high titers (.106 cfu/ml),36 and 3.75 mg of ZF-effector pMX retroviral vector using Lipofectamin PLUS transfection reagents as recommended by the manufacturer (Invitrogen). As mock infection control, the same infection conditions were used with the pcDNA3.1 vector (Invitrogen). Two days later, the viral supernatant was cleared by centrifugation and applied for eight hours with 8 mg/ml polybrene to 1 – 5 £ 105 host cells per 10 cm plate and infection was repeated overnight. The complexity of the 6ZF library (8.4 £ 107) necessitates a large number of infected cells, which was achieved by up-scaling the infection protocol. Infected cells were harvested for analysis 72 hours after infection. FACS analysis and cell sorting were done using Becton Dickinson flow cytometers. The anti-ICAM-1 primary antibody was purchased from BD PharMingen (clone HA58). Antibodies used for staining epidermal growth factor (EGF), 3-FAL selectin ligand (CD15, FUT4), integrin-a6 (CD49f, ITGA6), leukocyte function-associated antigen (CD58, LFA-3), Apo1-FAS antigen (CD95, TNFRSF6), integrin-b4 (CD104, ITGB4) and vascular endothelial VE-cadherin (CD144, CDH5) were used as reported.12 Comparative analysis of ICAM-1 flow cytometric data, mean fluorescence intensities (MFI), was determined using CELLQuest software (Becton Dickinson) as described.33 For each sample analyzed, the MFI was corrected for background by subtraction of the autofluorescence of cells incubated with the secondary antibody only and normalized to the value of cells infected with the unselected 6ZFlibrary and presented as fold activation or percentage remaining ICAM-1. For ICAM-1 induction studies, 20 ng/ml IL-1b (Sigma) and 10 ng/ml TNF-a (Sigma) were added to the cells 12 hours before the FACS analysis. CCATCTACAGCTT-30 ) and ICAMRT-1265-r reverse primers (50 -CAATCCCTCTCGTCCAGTCG-30 ); integrin alpha 6 (ITGA6) at 54 8C for 25 cycles using ITGA6-f1 (50 -AACTTGGACACTCGGGAGGACAAC-30 ) and ITGA6-b1 (50 -GGGGTCAGCATCGTTATCAAACTC-30 ) primers; glyceraldehyde-3-phosphate dehydrogenase (GAPDH) at 52 8C for 20 cycles with GAPDH forward and reverse primers,9 and TFZF-KRAB at 52 8C for 30 cycles with pMX-SKD-f (50 -ATCCGCCACCATGGATG30 ) and NLS-seq-b (50 -CTGGGCCACTTTGCGTTTC-30 ) primers, which amplify the 327 bp KRAB domain of the TFZF. The levels of mRNA expression were determined by quantifying the amount of PCR product on 1.5% agarose gel electrophoresis with ImageQuant software (Molecular Dynamics), background was subtracted and the signal normalized to GAPDH. The 303 bp ICAM-1 and 243 bp ITGA6 PCR products were confirmed by sequencing. Generation of CD54-31Opt protein The six-finger CD54-31Opt was assembled from two three-finger proteins constructed by PCR grafting of the appropriate DNA recognition helices into the framework of the three-finger protein Sp1C as described.6 The corresponding ZF domains and DNA subsites are shown in Table 1. Characterization of 6ZF proteins in vitro The selected and designed 6ZF protein coding regions were subcloned into Escherichia coli expression vector pMal-C2 (NEB) by SfII digestion. The purification of the resulting MBP fusion proteins and their DNA-binding properties were determined by multitarget specificity ELISA and electrophoretic mobility-shift assays (EMSA) and were carried out in duplicates essentially as described.3 For this purpose, biotinylated dsDNA oligonucleotides containing the 18 bp target sequences were synthesized (MWG-Biotech) and are presented in Table 1. DNA binding for ELISA is determined from serial dilutions in duplicate experiments, from which dilution series of 60 nM protein were used. The ICAM-1 promoter target oligonucleotides used by ELISA were named after their position relative to the start of translation. The DNA strand and the number of matching residues relative to the predicted 18 bp cd54-31 sequence are presented in parentheses. ICAM-1 pro-1511 (þ, 11) 50 -AGC TCAGTGGAACCCGCC-30 pro-940 (2 , 10) 50 -TGAGGA GTTCTGAATTCC-30 pro-1000 (þ, 11) 50 -GCAGGGGTT CGAGCGCCC-30 pro-793; (þ , 10) 50 -GGGAGCTGTAAA GACGCC-30 pro-757 (þ , 9) 50 -AGGGGAAGCGAGGA GGCC-30 pro-501 (þ, 11) 50 -CACTCGATTAAAGGGCC30 pro-220 (þ , 13) 50 -TCCGGAGCTGAAGCGGCC-30 pro-59 (- 2 , 11) 50 -TGATCCTTTATAGCGCTA-30 . Semi-quantitative RT-PCR DNase I footprinting Total RNAs were isolated from A431 cells using TRI reagent (Molecular Research Center), 1st strand cDNA was made using the Superscript reverse transcriptase (Invitrogen) as recommended by the manufacturer. Semi-quantitative RT-PCR was performed as described12 with the following modifications. In order to compare PCR products in the linear range of the PCR reaction, 2 ml of cDNA was used as template and ICAM-1 was amplified at a 54 8C annealing temperature for 25 cycles using the ICAMRT-963-f forward (50 -GCAGACAGTGA Binding of ZFPs to ICAM-1 DNA sequences was examined by DNase I footprinting by a following the method of Trauger & Dervan.37 The 200 bp DNA probe containing the ICAM-1 promoter site pro-220 was generated by PCR from the ICAM-1 promoter reporter construct as following. A forward DNA oligonucleotide primer 50 -GTCATCGCCCTGCCACCG-30 was labeled with polynucleotide kinase (Roche) and [g-32P]ATP (8000 Ci/mmol) (Perkin Elmer Life Sciences) as per manufacturer’s instruction. The radioactively labeled 648 primer was purified and used together with a reverse primer 50 -TTTATAGCGCTAGCCACCTGGGG-30 for PCR amplification with High Fidelity PCR Master mix (Roche) and 100 ng pGL3-ICAM-1 template plasmid for 35 amplification cycles (30 seconds at 94 8C, 30 seconds, at 55 8C, 30 seconds at 72 8C). The resulting DNA probe was recovered from a non-denaturing polyacrylamide gel and 17 – 24 ng was used in each DNase I protection assay. Binding equilibrations of ZFPs at 1 mM, 100 nM, 10 nM and 1 nM concentrations with DNA was done in 400 ml binding buffer (10 mM Tris – HCl (pH 7.0), 10 mM KCl, 10 mM MgCl2, 5 mM CaCl2, 10 mM ZnCl2 and 5 mM DTT) overnight at 4 8C. DNase I (RNase-free, Roche) was added to 3 mU/ml and digestion was stopped after seven minutes at room temperature. DNA was ethanol precipitated and electrophoretically separated on a 6% (w/v) polyacrylamide/8 M urea denaturing gel along with guanine-specific chemical cleavage products of the same DNA probe as marker.38 Transient reporter assay, ICAM-1 reporter plasmid and deletion mutant thereof A 1.3 kb upstream promoter fragment of the ICAM-1 gene was shown to drive luciferase expression and mediates responsiveness to ICAM-1 inducers.23 To generate the pGL3-ICAM-1 reporter construct, a 1.6 kb PCR fragment of the ICAM-1 promoter (nucleotide positions 2 1592 to 2 15 from the start of translation) was PCR amplified from HUVEC genomic DNA using the forward 50 -GAGGAGGAGGAGGAGGGTACCTGAGAAA AGAACGGCACCATTG-30 and reverse 50 -GAGGAGGA GGAGGAGACGCGTTGCAACTCTGAGTAGCAGAGG AGC-30 primers and was cloned into the pGL3-basic luciferase vector (Promega). The deletion mutant pGL3E2C/ICAM-1 was created by PCR mutagenesis. Transient reporter luciferase assays in A431 cells were done as reported.6 Luciferase activity was normalized to b-galactosidase activity using the Galacto-light Plus assay (Tropix). Data represent the average of three experiments. ZF antibody generation and ChIP assay Since at most seven out of 28 amino acid residues differ from one domain to another, a minimum of 75% conserved residues is provided overall between any designer zinc finger domain. Therefore, the polyclonal serum that was generated by rabbit immunization with a mixture of three purified 3ZF proteins encoding the consensus framework of Sp1C, was expected to recognize all Sp1C variants. ChIP assays were performed as described with the following modifications.39 Nuclei from a minimum of 107 cells per immunoprecipitation were cross-linked, isolated and sonicated to produce 500 bp chromatin fragments. TFZF : chromatin complexes were captured by centrifugation using 10 ml of ZF antibody or no antibody and lyophilized StaphA cells (Calbiochem), washed extensively and analyzed by PCR at a 60 8C annealing temperature for 40 cycles using the ICAMpro-1381-f1 GTACTTAATAAACCGATTAAGCG forward and ICAMpro-b5 TGCAACTCTGAGTAGCAG AGGAGC reverse primers producing a 343 bp PCR fragment containing the ICAM-1 pro-220/NF-kB target site. PCR reactions were performed with High fidelity Taq DNA polymerase (Roche) and 2 ml of IP sample or 0.2% of the total input chromatin that was not immunoprecipitated from the “no antibody” sample. Novel ICAM-1 Zinc Finger Transcription Factors Acknowledgements We thank Dave Valente for technical support. This study was supported in part by the National Institutes of Health grant CA086258 (to C.F.B.). L.M was the recipient of postdoctoral fellowships from the Swiss National Science Foundation. References 1. Greisman, H. A. & Pabo, C. O. (1997). A general strategy for selecting high-affinity zinc finger proteins for diverse DNA target sites. Science, 275, 657– 661. 2. Liu, Q., Segal, D. J., Ghiara, J. B. & Barbas, C. F., III (1997). Design of polydactyl zinc-finger proteins for unique addressing within complex genomes. Proc. Natl Acad. Sci. USA, 94, 5525– 5530. 3. Segal, D. J., Dreier, B., Beerli, R. R. & Barbas, C. F., III (1999). Toward controlling gene expression at will: selection and design of zinc finger domains recognizing each of the 50 -GNN-30 DNA target sequences. Proc. Natl Acad. Sci. USA, 96, 2758– 2763. 4. Dreier, B., Beerli, R. R., Segal, D. J., Flippin, J. D. & Barbas, C. F., III (2001). Development of zinc finger domains for recognition of the 50 -ANN-30 family of DNA sequences and their use in the construction of artificial transcription factors. J. Biol. Chem. 276, 29466– 29478. 5. Beerli, R. R. & Barbas, C. F., III (2002). Engineering polydactyl zinc-finger transcription factors. Nature Biotechnol. 20, 135– 141. 6. Beerli, R. R., Segal, D. J., Dreier, B. & Barbas, C. F., III (1998). Toward controlling gene expression at will: specific regulation of the erbB-2/HER-2 promoter by using polydactyl zinc finger proteins constructed from modular building blocks. Proc. Natl Acad. Sci. USA, 95, 14628– 14633. 7. Beerli, R. R., Dreier, B. & Barbas, C. F., III (2000). Positive and negative regulation of endogenous genes by designed transcription factors. Proc. Natl Acad. Sci. USA, 97, 1495– 1500. 8. Zhang, L., Spratt, S. K., Liu, Q., Johnstone, B., Qi, H., Raschke, E. E. et al. (2000). Synthetic zinc finger transcription factor action at an endogenous chromosomal site. Activation of the human erythropoietin gene. J. Biol. Chem. 275, 33850– 33860. 9. Liu, P. Q., Rebar, E. J., Zhang, L., Liu, Q., Jamieson, A. C., Liang, Y. et al. (2001). Regulation of an endogenous locus using a panel of designed zinc finger proteins targeted to accessible chromatin regions. Activation of vascular endothelial growth factor A. J. Biol. Chem. 276, 11323– 11334. 10. Rebar, E. J., Huang, Y., Hickey, R., Nath, A. K., Meoli, D., Nath, S. et al. (2002). Induction of angiogenesis in a mouse model using engineered transcription factors. Nature Med. 8, 1427– 1432. 11. Guan, X., Stege, J., Kim, M., Dahmani, Z., Fan, N., Heifetz, P. et al. (2002). Heritable endogenous gene regulation in plants with designed polydactyl zinc finger transcription factors. Proc. Natl Acad. Sci. USA, 99, 13296– 13301. 12. Blancafort, P., Magnenat, L. & Barbas, C. F. (2003). Scanning the human genome with combinatorial transcription factor libraries. Nature Biotechnol. 21, 269– 274. 649 Novel ICAM-1 Zinc Finger Transcription Factors 13. van de Stolpe, A. & van der Saag, P. T. (1996). Intercellular adhesion molecule-1. J. Mol. Med. 74, 13 – 33. 14. Sligh, J. E., Jr, Ballantyne, C. M., Rich, S. S., Hawkins, H. K., Smith, C. W., Bradley, A. & Beaudet, A. L. (1993). Inflammatory and immune responses are impaired in mice deficient in intercellular adhesion molecule 1. Proc. Natl Acad. Sci. USA, 90, 8529– 8533. 15. Kim, J. S. & Pabo, C. O. (1998). Getting a handhold on DNA: design of poly-zinc finger proteins with femtomolar dissociation constants. Proc. Natl Acad. Sci. USA, 95, 2812– 2817. 16. Dreier, B., Segal, D. J. & Barbas, C. F., III (2000). Insights into the molecular recognition of the 50 -GNN-30 family of DNA sequences by zinc finger domains. J. Mol. Biol. 303, 489– 502. 17. Segal, D. J. (2002). The use of zinc finger peptides to study the role of specific factor binding sites in the chromatin environment. Methods, 26, 76 –83. 18. Liu, X., Constantinescu, S. N., Sun, Y., Bogan, J. S., Hirsch, D., Weinberg, R. A. & Lodish, H. F. (2000). Generation of mammalian cells stably expressing multiple genes at predetermined levels. Anal. Biochem. 280, 20 – 28. 19. Margolin, J. F., Friedman, J. R., Meyer, W. K., Vissing, H., Thiesen, H. J. & Rauscher, F. J., III (1994). Kruppel-associated boxes are potent transcriptional repression domains. Proc. Natl Acad. Sci. USA, 91, 4509–4513. 20. Ayer, D. E., Laherty, C. D., Lawrence, Q. A., Armstrong, A. P. & Eisenman, R. N. (1996). Mad proteins contain a dominant transcription repression domain. Mol. Cell. Biol. 16, 5772– 5781. 21. Jiang, X. R., Jimenez, G., Chang, E., Frolkis, M., Kusler, B., Sage, M. et al. (1999). Telomerase expression in human somatic cells does not induce changes associated with a transformed phenotype. Nature Genet. 21, 111 – 114. 22. Morales, C. P., Holt, S. E., Ouellette, M., Kaur, K. J., Yan, Y., Wilson, K. S. et al. (1999). Absence of cancerassociated changes in human fibroblasts immortalized with telomerase. Nature Genet. 21, 115 – 118. 23. Voraberger, G., Schafer, R. & Stratowa, C. (1991). Cloning of the human gene for intercellular adhesion molecule 1 and analysis of its 50 -regulatory region. Induction by cytokines and phorbol ester. J. Immunol. 147, 2777– 2786. 24. Krizek, B. A., Amann, B. T., Kilfoil, V. J., Merkle, D. L. & Berg, J. M. (1991). A consensus zinc finger peptide—design, high-affinity metal-binding, a Ph-dependent structure, and a His to Cys sequence variant. J. Am. Chem. Soc. 113, 4518– 4523. 25. Shi, Y. & Berg, J. M. (1995). A direct comparison of the properties of natural and designed zinc-finger proteins. Chem. Biol. 2, 83 – 89. 26. Ledebur, H. C. & Parks, T. P. (1995). Transcriptional regulation of the intercellular adhesion molecule-1 gene by inflammatory cytokines in human endothelial cells. Essential roles of a variant NF-kappa B site and p65 homodimers. J. Biol. Chem. 270, 933– 943. 27. Liang, Y., Li, X. Y., Rebar, E. J., Li, P., Zhou, Y., Chen, B. et al. (2002). Activation of vascular endothelial growth factor A transcription in tumorigenic 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. glioblastoma cell lines by an enhancer with cell type-specific DNase I accessibility. J. Biol. Chem. 277, 20087 – 20094. Roebuck, K. A. & Finnegan, A. (1999). Regulation of intercellular adhesion molecule-1 (CD54) gene expression. J. Leukoc. Biol. 66, 876– 888. Segal, D. J., Beerli, R. R., Blancafort, P., Dreier, B., Effertz, K., Huber, A. et al. (2003). Evaluation of a modular strategy for the construction of novel polydactyl zinc finger DNA-binding proteins. Biochemistry, 42, 2137– 2148. van de Stolpe, A., Caldenhoven, E., Stade, B. G., Koenderman, L., Raaijmakers, J. A., Johnson, J. P. & van der Saag, P. T. (1994). 12-O-tetradecanoylphorbol-13-acetate- and tumor necrosis factor alphamediated induction of intercellular adhesion molecule-1 is inhibited by dexamethasone. Functional analysis of the human intercellular adhesion molecular-1 promoter. J. Biol. Chem. 269, 6185– 6192. Mulder, W. M., Stern, P. L., Stukart, M. J., de Windt, E., Butzelaar, R. M., Meijer, S. et al. (1997). Low intercellular adhesion molecule 1 and high 5T4 expression on tumor cells correlate with reduced disease-free survival in colorectal carcinoma patients. Clin. Cancer Res. 3, 1923– 1930. Ogawa, Y., Hirakawa, K., Nakata, B., Fujihara, T., Sawada, T., Kato, Y. et al. (1998). Expression of intercellular adhesion molecule-1 in invasive breast cancer reflects low growth potential, negative lymph node involvement, and good prognosis. Clin. Cancer Res. 4, 31 – 36. Griffioen, A. W., Damen, C. A., Mayo, K. H., Barendsz-Janson, A. F., Martinotti, S., Blijham, G. H. & Groenewegen, G. (1999). Angiogenesis inhibitors overcome tumor induced endothelial cell anergy. Int. J. Cancer, 80, 315– 319. Johnson, J. P., Stade, B. G., Holzmann, B., Schwable, W. & Riethmuller, G. (1989). De novo expression of intercellular-adhesion molecule 1 in melanoma correlates with increased risk of metastasis. Proc. Natl Acad. Sci. USA, 86, 641– 644. Naldini, L., Blomer, U., Gallay, P., Ory, D., Mulligan, R., Gage, F. H. et al. (1996). In vivo gene delivery and stable transduction of nondividing cells by a lentiviral vector. Science, 272, 263– 267. Burns, J. C., Friedmann, T., Driever, W., Burrascano, M. & Yee, J. K. (1993). Vesicular stomatitis virus G glycoprotein pseudotyped retroviral vectors: concentration to very high titer and efficient gene transfer into mammalian and nonmammalian cells. Proc. Natl Acad. Sci. USA, 90, 8033– 8037. Trauger, J. W. & Dervan, P. B. (2001). Footprinting methods for analysis of pyrroleimidazole polyamide/DNA complexes. Methods Enzymol. 340, 450 –466. Maxam, A. M. & Gilbert, W. (1980). Sequencing endlabeled DNA with base-specific chemical cleavages. Methods Enzymol. 65, 499– 560. Boyd, K. E., Wells, J., Gutman, J., Bartley, S. M. & Farnham, P. J. (1998). c-Myc target gene specificity is determined by a post-DNA binding mechanism. Proc. Natl Acad. Sci. USA, 95, 13887– 13892. Edited by M. Yaniv (Received 19 January 2004; received in revised form 11 June 2004; accepted 14 June 2004)