allen et al., 2007

Upload: danny-allen

Post on 08-Apr-2018

231 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/6/2019 Allen et al., 2007

    1/12

    THE JOURNAL OF GENE MEDICINE R E S E A R C H A R T I C L EJ Gene Med 2007; 9: 287298.Published online in Wiley InterScience (www.interscience.wiley.com) DOI: 10.1002/jgm.1018

    Development of strategies for conditional RNAinterference

    Danny Allen*

    Paul F. Kenna

    Arpad Palfi

    Helena P. McMahon

    Sophia Millington-WardMary OReilly

    Pete Humphries

    G. Jane Farrar

    Department of Genetics, Trinity

    College Dublin, Dublin 2, Ireland

    *Correspondence to: Danny Allen,

    Department of Genetics, Trinity

    College Dublin, Dublin 2, Ireland.

    E-mail: [email protected]

    Received: 19 September 2006

    Revised: 22 December 2006

    Accepted: 22 January 2007

    Abstract

    Background RNA interference (RNAi) represents a powerful tool with

    which to undertake sequence-dependent suppression of gene expression.

    Synthesized double-stranded RNA (dsRNA) or dsRNA generated endoge-nously from plasmid or viral vectors can be used for RNAi. For the latter,

    polymerase III promoters which drive ubiquitous expression in all tissues

    have typically been adopted. Given that dsRNA molecules must contain few

    5 and 3 over-hanging bases to maintain potency, employing polymerase II

    promoters to drive tissue-specific expression of RNAi may be problematic due

    to potential inclusion of nucleotides 5 and 3 of siRNA sequences.

    Methods To circumvent this, polymerase II promoters in combination with

    cis-acting hammerhead ribozymes and short-hairpin RNA sequences have

    been explored as a means to generate potent dsRNA molecules in tissues

    defined by the promoter in use.

    Results The novel constructs evaluated in this study produced functional

    siRNA which suppressed the enhanced green fluorescent protein (eGFP) both

    in vitro and in vivo (in mice). Additionally, the constructs did not appear

    to elicit a significant type-1 interferon response compared to traditional

    H1-transcribed shRNA.

    Conclusions Given the potential off-target effects of dsRNAs, it would

    be preferable in many cases to limit expression of dsRNA to the tissue of

    interest and moreover would significantly augment the resolution of RNAi

    technologies. Notably, the system under evaluation in this study could readily

    be adapted to achieve this objective. Copyright 2007 John Wiley & Sons,

    Ltd.

    Keywords RNA interference; ribozymes; polymerase II promoters; gene therapy

    Introduction

    Use of double-stranded RNA (dsRNA) to modulate gene expression, or

    RNA interference (RNAi), has been adopted extensively subsequent to the

    breakthrough paper of Andrew Fire and colleagues in 1998 demonstrat-

    ing potent sequence-specific gene silencing in Caenorhabditis elegans [1].

    Observation of an interferon response in mammalian cells to long dsRNA

    molecules precipitated the use of short dsRNA molecules or small inter-

    fering RNA (siRNA) to circumvent this response [2]. Much of the initialresearch was focused on exploration of the utility of synthesized siRNAs,

    which elicit transient suppression of a target gene. The transience of siRNA-

    based suppression promoted the emergence of chemically protected siRNA

    Copyright 2007 John Wiley & Sons, Ltd.

  • 8/6/2019 Allen et al., 2007

    2/12

    288 D. Allen et al.

    molecules incorporating modifications thereby providing

    molecules with longer half-lives [3].

    An alternative means of overcoming transient suppres-

    sion involves engineering vector(s) from which functional

    siRNAs may be expressed [4 9]. Short hairpin RNAs

    (shRNAs) expressed from such vectors are processed

    intracellularly into functional siRNAs [10]. Expression

    of shRNAs can be achieved using various systems; e.g.

    ubiquitously expressing polymerase III promoters such

    as the H1 or U6 promoters which utilize short initia-

    tion and termination signals have been employed [48].

    However, the majority of mammalian promoters are poly-

    merase II promoters and typically use more complex

    initiation and termination signals than their polymerase

    III counterparts. The cytomegalovirus (CMV) promoter, a

    ubiquitous polymerase II promoter, has been employed to

    drive expression of shRNAs; typically, shRNA sequences to

    be expressed have been placed juxtaposed to the transcrip-

    tional start site [11]. In contrast to CMV, transcriptionalstart sites for many polymerase II promoters are ill defined

    or multiple transcriptional start sites may be utilized.

    Nevertheless, there are clear advantages to targeting

    RNAi-based suppression to specific tissues particularly

    in the light of potential off-target effects associated

    with RNAi-based suppression identified from microarray

    studies [12]. Additionally, such tissue specificity would

    increase the resolution of RNAi technology, in principle,

    mirroring the resolution achieved with conditional gene

    targeting.

    Materials and methods

    Construction of RNAi plasmids

    Polymerase chain reactions (PCRs) were carried out

    under the following conditions. The PCR reaction had

    a final volume of 20 l containing 5 pmol of forward

    and reverse primer, 0.5pmol of forward and reverse

    template oligonucleotide, 0.2 M dATP, dTTP, dCTP, and

    dGTP, 5 l 10 buffer and 1 unit of Taq polymerase

    (10 buffer = 500 mM KCl, 100 mM Tris, pH 9, 0.1%

    gelatin (v/v), 1% Triton X and 15 mM MgCl2

    ). The

    standard PCR cycle run was; 94 C for 5 min followed

    by 35 cycles of 94 C for 1 min, 55 C for 1 min and 72 C

    for 1.2 min, ending with a final step of 72 C for 10 min.

    All primer and template oligonucleotide sequences are

    detailed in Table 1. PCR products were cloned into

    pCDNA3.1 (Invitrogen) using restriction enzyme sites

    incorporated into the forward and reverse primers. Colony

    PCR screens were carried out to detect positive colonies.

    DNA from positive colonies was sequenced to ensure

    synthesis fidelity of the cloned PCR products.

    Cell transfection and RNA extraction

    HeLa cells were cultured in DMEM+ (500 ml Dulbeccos

    modified Eagles medium (DMEM), 50 ml foetal calf

    serum (FCS), 5 ml 100 mM sodium pyruvate, 5 ml

    L-glutamine and 5 ml penicillin/streptomycin) using

    standard procedures. All co-transfections (enhanced

    green fluorescent protein (eGFP) plasmid + test/control

    plasmid) were carried out in triplicate in six-well plates.

    Twenty-four hours prior to transfection, 5 105 cells

    were plated in each well of a six-well plate and

    grown under standard conditions minus antibiotics.

    Transfections were carried out using Lipofectamine 2000

    as per the manufacturers instructions (Invitrogen).

    RNA extraction

    Twenty-four hours post-transfection total RNA was

    isolated from cells using TRI reagent (MRC) as per the

    manufacturers instructions. RNA was then treated with

    RNase-free DNase (Promega). RNAs extracted from HeLa

    cells were resuspended in 100 l of nuclease-free water

    and stored at 80 C. When enriching for small RNAs the

    mirVana miRNA isolation kit was used, instead of TRI

    reagent, as per the manufacturers instructions (Ambion).

    RNA extraction and real-timereverse-transcription (rt)PCR

    RNA levels were analyzed using real-time rtPCR carried

    out on a Lightcycler (Roche Diagnostics) using the

    Table 1. Primer and template oligonucleotide sequences used in this study

    tsRNAi Primer Forward CTAGCTAGCTCTAGAGGAtsRNAi Primer Reverse CCCAAGCTTGAATTCCACtsRNAi A eGFP Oligo CTAGCTAGCTCTAGAGGATCCAGTAGCTGATGAGTCCGTGAGGACGAAACGGTACCCGGTAGCAAGCTGACCCTGAAGTTCATCAGA

    AGAGAACTTCAGGGTCAGCTTGCCGGTCGACGGATCATGATCCGTCCTGATGAGTCCGTGAGGACGAAACAACGAATTCAAGCTTtsRNAi B eGFP Oligo CCTAGCGGATCCAGTAGCTGATGAGTCCGTGAGGACGAAACGGTACCCGGTACCGTCGCAAGCTGACCCTGAAGTTCATGACGGATC

    TAGATCCGTCCTGATGAGTCCGTGAGGACGAAACTGGGTCGCTAAAGCCCACCCAGCTGATGAGTCCGTGAGGACGAAACGGTACCGTCGAACTTCAGGGTCAGCTTGCCGGACGGATCATGATCCGTCCTGATGAGTCCGTGAGGACGAAACAACCACCAACGAATTCAAGCTT

    tsRNAi C eGFP Oligo CCTAGCGGATCCAGTAGCTGATGAGTCCGTGAGGACGAAACGGTACCCGGTACCGTCCTACTGCAAGCTGACCCTGAAGTTCATCAGAAGAGAACTTCAGGGTCAGCTTGCCGAATAAAGGAAATTTATTTTCATTGCAATAGTGTGTTGGTTTTTTGTGTGAAGCTTCCTAGC

    tsRNAi A non Oligo CTAGCTAGCTCTAGAGGATCCAGTAGCTGATGAGTCCGTGAGGACGAAACGGTACCCGGTATTCTCCGAACGTGTCACGTTTCAGAAGAACGTGACACGTTCGGAGAATTGTCGACGGATCATGATCCGTCCTGATGAGTCCGTGAGGACGAAACAACGAATTCAAGCTT

    tsRNAi B non Oligo CCTAGCGGATCCAGTAGCTGATGAGTCCGTGAGGACGAAACGGTACCCGGTACCGTCTTCTCCGAACGTGTCACGTTTGACGGATCTA

    GATCCGTCCTGATGAGTCCGTGAGGACGAAACTGGGTCGCTAAAGCCCACCCAGCTGATGAGTCCGTGAGGACGAAACGGTACCGTCACGTGACACGTTCGGAGAATTGACGGATCATGATCCGTCCTGATGAGTCCGTGAGGACGAAACAACCACCAACGAATTCAAGCTT

    tsRNAi C non Oligo CCTAGCGGATCCAGTAGCTGATGAGTCCGTGAGGACGAAACGGTACCCGGTACCGTCCTACTTTCTCCGAACGTGTCACGTTTCAGAAGAACGTGACACGTTCGGAGAATTAATAAAGGAAATTTATTTTCATTGCAATAGTGTGTTGGTTTTTTGTGTGAAGCTTCCTAGC

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    3/12

    Polymerase II Promoter-Driven RNAi 289

    Table 2. Real-time rtPCR primer sequences

    eGFP Forward TTCAAGGAGGACGGCAACATCCeGFP Reverse CACCTTGATGCCGTTCTTCTGCActin Forward TCACCCACACTGTGCCCATCTACGAActin Reverse CAGCGGAACCGCTCATTGCCAATGGGAPDH Forward GAAGGTGAAGGTCGGAGTC

    GAPDH Reverse GAAGATGGTGATGGGATTTCPPIA Forward CCCACCGTGTTCTTCGACATPPIA Reverse CCAGTGCTCAGAGCACGAAAOAS1 Reverse CAGCTTCGTACTGAGTTCGCOAS1 Reverse TAGTTCTGTGAAGCAGGTGGIFITM1 Forward CAACACCCTCTTCTTGAACTGGIFITM1 Reverse AGATGTTCAGGCACTTGGCGISGF3 Forward AAGTACCATCAAAGCGACAGCISGF3 Reverse CATTATTGAGGGAGTCCTGGEIF2AK2 Forward ACCTCAGTGAAATCTGACTACCEIF2AK2 Reverse CAGATGATGATTCAGAAGCG

    QuantiTect SYBR Green RT-PCR kit (Qiagen). Samples

    were run under the following conditions; an rt step (50C,

    20 min), followed by an inactivation step (95

    C, 15 min)and 35 cycles of an amplification step (94 C, 15 s;

    55 C, 20 s; 72 C, 4 s). Mean values, standard deviations

    and t-tests were calculated using DataDesk v6.0 (see

    below). Differences in levels of expression observed

    between samples were deemed statistically significant

    at p < 0.05. All real-time rtPCR primer sequences are

    detailed in Table 2. The cycle-threshold method of real-

    time rtPCR analysis was used for both mRNA relative

    quantification and interferon-related gene expression. The

    2CT method of determining mean fold changes in gene

    expression was used to analyze the interferon-related

    gene expression.

    Hydrodynamic tail-vein injection

    Endotoxin-free plasmid DNA was diluted in sterile

    phosphate-buffered saline (PBS). Total injection volumes

    per mouse (in ml) were calculated by dividing the weight

    of the mouse (in g) by 10. Unanesthetized mice were

    restrained in a 50 ml syringe barrel with rubber bungs.

    Once in position, a lamp was used to warm the mouse-tail

    and thereby dilate the tail vein facilitating visualization

    and injection with a 3-ml syringe, fitted with a 0.75-cm

    26-gauge needle. The syringe needle was placed into the

    dilated tail vein and nearly the full length of the needle

    inserted into the vein. Once inserted, the complete volume

    of solution was injected into the tail vein within 35 s.

    Notably, hydrodynamic injection appears to result in a

    temporary minor heart failure in the mouse, resulting in

    retrograde blood flow which pushes the injected sample

    back into the liver.

    Direct inferior vena cava injection

    Endotoxin-free plasmid DNA was diluted in sterile

    Hartmanns solution. Total injection volume per mouse was 500l. Mice were anesthetized prior to injection

    using ketamine/xylazine. In anesthetized mice the inferior

    vena cava (IVC) was exposed using a midline incision. A

    1-ml syringe, fitted with a 30-gauge needle, was used

    to deliver the DNA solution. Once the needle had been

    inserted, the complete volume of solution was injected

    into the IVC within 10 s.

    RNA extraction from mouse organs

    Forty-eight hours post injection, mice were sacrificed

    using CO2 asphyxiation. Organs collected from mice were

    homogenized in 1 ml of TRI reagent (MRC) using sterile

    polypropylene pellet pestles (Sigma Aldrich). Total RNA

    was isolated as per the manufacturers instructions. RNA

    was then treated with RNase-free DNase (Promega). RNA

    extracted from mouse liver was resuspended in 500 l of

    nuclease-free water and stored at 80 C.

    Tissue fixation and sectioningOrgans harvested for sectioning were preserved in 4%

    paraformaldehyde and stored at 4 C in the dark until

    embedded in 6% agar for sectioning. Sections of 100 m

    thickness were cut on a vibratome (Leica VT1000s) at

    a blade vibrating frequency of 70 Hz and a speed of

    0.75 mm/s. Analysis of sections was performed with a

    fluorescence microscope (Axiophot; Zeiss Ltd., UK) at

    12.5 magnification.

    Statistical analysis

    Statistics were performed using DataDesk v6.0.

    Unpaired two-sample t-tests were performed on the data

    to compare the differences between sample means.

    Animal handling

    All animals used in these experiments complied with the

    ARVO statement for the Use of Animals in Ophthalmic and

    Vision Research and were also passed by an intuitional

    ethics committee and under licence. Mice were held in the

    Barrier Mouse Facility in the Smurfit Institute of Genetics

    prior and subsequent to experimental procedures. All micewere sacrificed by CO2 asphyxiation.

    Results

    Expression cassettes have been generated in which a

    combination of polymerase II promoters in conjunction

    with cis-acting hammerhead ribozymes have been used

    to express functional shRNAs targeting eGFP and eGFP

    suppression; these have been evaluated initially in cell

    culture and subsequently in mice. Three constructs

    A C with unifying features, most notably cis-actinghammerhead ribozymes, have been explored. In essence,

    cis-acting ribozymes are employed to cleave nucleotides

    5 and 3 of antisense and sense sequences comprising

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    4/12

    290 D. Allen et al.

    dsRNAs thereby promoting generation of functional

    siRNAs. In each case, the sequence of one arm of the

    cis-acting ribozyme is in part complimentary to either

    the sense or antisense sequence used to form siRNAs.

    Construct A comprises an shRNA sequence cleaved 5

    and 3 by cis-acting hammerhead ribozymes. In contrast,

    four cis-acting hammerhead ribozymes are employed in

    construct B to generate sense and antisense strands

    separately which must hybridize within the cell to

    produce functional siRNAs. Construct C comprises a

    single hammerhead ribozyme 5 of the sense portion

    of the shRNA and a minimal poly-A signal at the

    3 end placed juxtaposed to sequence coding for the

    antisense portion of the shRNA. In all three eGFP targeting

    constructs (AC) the shRNA target sequence was

    CGGCAAGCTGACCCTGAAGTTCAT and for the three non-

    targeting constructs (A C) the shRNA target sequence

    was AATTCTCCGAAC GTGTCACGT. Both target sites are

    based on Qiagen control siRNA target sites with thenon-targeting shRNA sequence having no homology to

    any known mammalian genes. Both the eGFP-targeting

    constructs and the non-targeting constructs were placed

    under the control of a cytomegalovirus (CMV) promoter

    (Figures 1A1C).

    In principle, upon expression of the CMV-driven AC

    constructs, the cis-acting hammerhead ribozymes should

    self-cleave releasing shRNAs in the case of constructs

    A and C and sense and antisense RNA in the case of

    construct B. Given that the constructs described above

    comprise multiple elements, to ensure that constructs

    A C were indeed capable of expressing siRNAs, HeLa cells were transfected with these constructs. Total RNA was

    subsequently extracted and enriched for small RNAs using

    the mirVana miRNA isolation kit. RNA concentrations

    were calculated by absorbance at 260 nm and RNA purity

    analyzed using the A260 to A280 ratio. RNA preparations,

    enriched for siRNAs, were run on 4% TBE NuSieve3 : 1

    agarose gels using a dsRNA ladder (New England

    Biolabs) to size RNAs. RNA bands in these samples were

    compared to bands present in cells transfected with either

    the positive control (H1-shRNA(eGFP)), empty vector

    (pCDNA3.1) or untransfected cells. As shown in Figure 2,

    the results suggest that constructs AC and the positive

    control are all capable of producing siRNA as suggestedby the presence of bands of the correct size in the

    RNA samples from these transfections. In contrast no

    such bands were observed in the negative controls. It is

    notable that whilst design B produces separate sense and

    antisense strands of siRNA that have to anneal in vitro

    to form siRNAs, it appears that this does indeed occur

    (Figure 2).

    To examine whether the siRNAs being produced by con-

    structs AC were indeed functional, suppression studies

    were undertaken in HeLa cells using a CMV promoter

    to drive expression of the sequences. To determine

    the optimal experimental conditions, HeLa cells wereco-transfected with an eGFP reporter plasmid (pEGFP-

    C1, Clontech) and varying concentrations of the design

    C expression plasmid (between 0.1 and 10 g). eGFP

    transcript levels were measured by real-time rtPCR,

    normalized using glyceraldehyde-3-phosphate dehydro-

    genase (GAPDH), actin and cyclophilin-A expression

    and compared to eGFP expression in cells transfected

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    5/12

    Polymerase II Promoter-Driven RNAi 291

    with the non-targeting design C construct. As shown in

    Figure 3, the change in eGFP transcript level was sig-

    nificant (p < 0.05) using 1 g of plasmid resulting in

    a 53% decrease in expression. At 5 and 10 g of plas-

    mid the fold change was also significant (p < 0.05) with

    a 73% and 74% decrease in expression, respectively.Constructs AC were then co-transfected with an eGFP

    reporter plasmid into HeLa cells as above (5 g of con-

    struct with 1 g of eGFP reporter plasmid) (Figure 4).

    Constructs AC resulted in approximately 71%, 70% and

    73% suppression of eGFP expression, respectively, while

    the H1-shRNA(eGFP) positive control resulted in 58%

    suppression of the target (when compared to the appro-

    priate non-targeting shRNA control). The low level of

    suppression with the H1-shRNA(eGFP) may possibly be

    attributed to co-transfection of the two plasmids the

    target and the suppressor plasmids. Additionally, these

    constructs are driven by different promoters, i.e. a CMV-

    driven eGFP reporter and a H1-driven shRNA. The data

    provide a clear indication that functional siRNAs can be

    generated by exploiting hammerhead ribozymes to trim 5

    and 3 sequences flanking core elements (sense/antisense

    sequences) comprising a dsRNA molecule. Indeed sup-

    pression of the eGFP target was achieved by either

    expressing shRNA sequences (constructs A and C) or

    sense and antisense siRNA sequence (construct B). In prin-

    ciple, the same approach could be utilized subsequently

    incorporating tissue-specific polymerase II promoters to

    provide tissue specificity of RNAi. Trimming of sequences

    5 and 3 of the RNAi molecule allows for flexibility in

    choice of polymerase II promoter and hence would enabletissue-specific RNAi.

    Figure 1. Diagrammatic representation of constructs AC.

    (A) Construct A utilizes two cis-acting hammerhead ribozymes

    to cleave 5 and 3 of an shRNA sequence. (B) In contrast,

    construct B utilizes four cis-acting hammerhead ribozymes,

    two that cleave 5 and 3 of the siRNA sense strand and two

    that cleave 5 and 3 of the siRNA antisense strand. Sense

    and antisense strands must anneal intracellularly to generate

    functional siRNAs. (C) Construct C utilizes a single cis-acting

    ribozyme to cleave 5 of the shRNA sequence with a minimal

    polyadenylation signal immediately 3 of the shRNA sequence.

    All three constructs are expressed using a CMV promoter.

    Additionally, constructs AC have been generated carryingeither an shRNA sequence targeting eGFP or a non-targeting

    control shRNA sequence

    In some instances siRNAs and shRNAs have been found

    to elicit a type-1 interferon response. It was important

    to determine if the presence of additional elements in

    the constructs, e.g. cis-acting hammerhead ribozymes,

    may potentially result in an increased type-1 interferon

    response, which clearly would not be desirable. Todetermine whether constructs A C stimulate a type-1

    interferon response the expression of several interferon-

    stimulated genes (OAS1, IFITM1, ISGF3g and EIF2AK2)

    was compared to their expression levels in cells co-

    transfected with either the H1-shRNA(eGFP) positive

    control (directed towards the same eGFP target) or

    the non-targeting shRNA control. Transcript levels were

    measured by real-time rtPCR and normalized using

    GAPDH, actin and cyclophilin-A expression. For the

    interferon response the 2CT method was used to

    determine mean fold changes in gene expression [13].

    As shown in Figure 5, the resulting data indicated that

    there was no significant change in expression of IFITM1

    (p = 0.56) and EIF2AK2 (p = 0.1) with constructs AC

    when compared to the non-targeting control construct.

    Although OAS1 expression varied, e.g. a 2.3-fold increase

    in OAS1 expression was observed with construct A, there

    was no significant (p = 0.25) change in OAS1 expression

    compared to the H1-shRNA(eGFP) positive control.

    Furthermore, ISGF3g expression increased between 2.3-

    and 3.5-fold with constructs A C; however, this increase

    in ISGF3g expression was not significant (p = 0.24)

    compared to the H1-shRNA(eGFP) positive control. In

    conclusion, the results obtained suggest that, despite

    the inclusion of additional components in constructs A C such as cis-acting hammerhead ribozymes, these

    constructs operate in a similar manner to standard H1-

    driven shRNA constructs in terms of the immune response

    they elicit.

    Given that from the in vitro work described above

    it is clear that cis-acting ribozymes can act in concert

    with shRNA sequences to produce functional siRNAs

    and thereby elicit suppression of a target gene in cell

    culture, it was timely to explore if this strategy would

    also be effective in vivo. Hence eGFP-expressing mice,

    C57BL/6-Tg(ACTB-eGFP)1Osb/J (Jackson Laboratories),

    were used to explore if functional siRNAs couldbe generated from constructs A C in vivo. C57BL/6-

    Tg(ACTB-eGFP)1Osb/J mice express an eGFP gene driven

    Figure 2. The four top panels depict photographs of 4% TBE NuSieve3 : 1 agarose gels (under UV light) with a dsRNA ladder and

    approx. 2 g of RNA enriched for small RNAs from HeLa cells transfected with the empty vector (pCDNA3.1) (A), positive control

    (H1-shRNA(eGFP)) (B), construct A (C), construct B (D), or construct C (E)

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    6/12

    292 D. Allen et al.

    Figure 3. Graph of eGFP transcript levels in HeLa cells co-transfected with construct C at various concentrations and an eGFP

    reporter plasmid. To determine the optimal plasmid concentration for further experiments, increasing quantities of construct Cwere co-transfected with an eGFP reporter plasmid. eGFP transcript levels were measured by real-time rtPCR and normalized using

    GAPDH, actin and cyclophilin-A expression. The eGFP expression obtained was compared to eGFP expression in cells transfected

    with non-targeting constructs. The reduction in eGFP transcript levels was significant (p < 0.05) using 1 g of the construct C

    plasmid (53% decrease in mRNA levels). At 5 g and 10 g of plasmid the fold change was also significant (p < 0.05) with a 73%

    and 74% decrease in eGFP mRNA, respectively. Error bars represent standard deviation

    Figure 4. Graph of eGFP transcript levels in HeLa cells co-transfected with constructs A C and an eGFP reporter plasmid. eGFP

    transcript levels were measured by real-time rtPCR and normalized using GAPDH, actin and cyclophilin-A expression. Expression

    was compared to eGFP expression in cells transfected with non-targeting constructs. Cells transfected with the positive control

    construct (H1-shRNA(eGFP)) showed 58% suppression of eGFP transcript levels. Cells transfected with constructs A, B and C

    showed 71%, 70% and 73% suppression of eGFP transcript levels, respectively. Notably, all four eGFP-targeting constructs resulted

    in significant eGFP suppression when compared to the appropriate non-targeting shRNA control construct (p < 0.05). Error bars

    represent standard deviation

    by a chicken beta-actin promoter with a CMV enhancer

    [14]. Sixty-four transgenic mice were treated twice over

    a 24-h period with 20 g of constructs AC (either the

    eGFP-targeting or the non-targeting versions), the H1-

    shRNA(eGFP) positive control or the non-targeting H1-shRNA, administered via hydrodynamic tail-vein injection

    (total 40 g). Liver, kidney and spleen were harvested

    48 h after the second injection. RNA extracted from tissues

    was analyzed for eGFP expression by real-time rtPCR and

    normalized using GAPDH and ribosomal 18s expression.

    eGFP expression in mice injected with eGFP-targeting

    constructs AC was compared to eGFP expression in mice

    injected with the non-targeting control constructs AC.Results obtained in vivo were similar to those observed

    in the experiments undertaken in cell culture (Figure 6).

    In liver, the positive control construct H1-shRNA(eGFP)

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    7/12

    Polymerase II Promoter-Driven RNAi 293

    Figure 5. Graphs of OAS1 (A), IFITM1 (B), ISGF3g (C), and EIF2AK2 (D) transcript levels in HeLa cells co-transfected with constructs

    AC, the H1-shRNA(eGFP) or the non-targeting shRNA control and an eGFP reporter plasmid. Transcript levels were measured by

    real-time rtPCR and normalized using GAPDH, actin and cyclophilin-A expression. Expression was compared to cells transfected

    with either the H1-shRNA(eGFP) positive control (directed towards the same eGFP target) or the non-targeting shRNA control.

    The resulting data indicated no significant change in expression of IFITM1 (p = 0.56) and EIF2AK2 (p = 0.1) compared to the

    non-targeting control. Although OAS1 expression varied, e.g. a 2.3-fold increase in OAS1 expression was observed with design A,

    there was no significant (p = 0.25) change in OAS1 expression compared to the H1-shRNA(eGFP) positive control. Furthermore,ISGF3g expression increased between 2.3- and 3.5-fold with constructs AC; however, this increase in ISGF3g expression was not

    significant (p = 0.24) compared to the H1-shRNA(eGFP) positive control. Error bars represent standard deviation

    resulted in a 45% decrease in eGFP transcript levels.

    Similarily, constructs A C resulted in 43%, 40% and

    54% decreases in eGFP transcript levels respectively in

    liver. Notably, eGFP suppression with constructs A C

    in liver did not differ significantly from the positive

    control (p = 0.57). In contrast, all four eGFP-targeting

    RNAi constructs demonstrated a significant difference, in

    terms of eGFP suppression, from the non-targeting shRNA

    negative control construct (p < 0.05). Furthermore, inaddition to liver, eGFP suppression was evaluated in

    kidney and spleen. Whilst no eGFP suppression was seen

    in kidney or spleen it was established, via the injection

    of an eGFP reporter plasmid, that hydrodynamic injection

    resulted in preferential delivery to liver with no significant

    expression in other organs such as kidney and spleen

    (discussed below).

    In addition to evaluation using real-time rtPCR,

    liver samples from all 64 injected mice were also

    taken and preserved in paraformaldehyde for fluorescent

    microscopy. The four panels shown in Figure 7A depict

    representative 100-m vibratome liver sections (12.5

    ),under UV light and a GFP filter, from C57BL/6-Tg(ACTB-

    EGFP)1Osb/J mice tail vein injected with a total of

    40 g of a non-targeting H1-shRNA and non-targeting

    constructs A C (A D, respectively). A significant level

    of fluorescence can be observed in each of the sections.

    In these sections it is possible to view an entire liver

    cross-section demonstrating even distribution of eGFP

    fluorescence (Figure 7A). In contrast, the four panels in

    Figure 7B depict representative 100-m vibratome liver

    sections (12.5), under UV light and a GFP filter, from

    C57BL/6-Tg(ACTB-EGFP)1Osb/J mice tail vein injected

    with a total of 40g of eGPF targeting H1-shRNAand eGFP-targeting constructs AC (AD, respectively).

    Notably, a significant reduction in green fluorescence

    can be observed in each of the sections in Figure 7B

    compared to those shown in Figure 7A. In these sections it

    is possible to view an entire liver cross-section in this case

    demonstrating the even distribution of eGFP suppression

    (Figure 7B). In summary, the results from in vivo delivery

    of constructs AC in mice mirror those obtained in cell

    culture and suggest that these constructs incorporating

    cis-acting hammerhead ribozymes and shRNA sequences

    may be used to generate potent siRNAs in vivo from

    polymerase II promoters.To determine whether constructs A C stimulate a

    type-1 interferon response in vivo the expression of several

    interferon-stimulated genes (OAS1, IFITM1, ISGF3g and

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    8/12

    294 D. Allen et al.

    Figure 6. Graph of eGFP transcript levels in liver, kidney and spleen from C57BL/6-Tg(ACTB-EGFP)1Osb/J mice. The

    H1-shRNA(eGFP) positive control, and both eGFP-targeting and non-targeting constructs AC, were evaluated in vivo subsequent to

    hydrodynamic tail-vein injection of C57BL/6-Tg(ACTB-EGFP)1Osb/J mice. A total of 64 mice were treated twice over a 24-h period

    with 20 g of plasmid (total of 40 g). eGFP transcript levels were measured by real-time rtPCR and normalized using GAPDH and

    ribosomal 18s expression. eGFP expression in mice injected with eGFP-targeting constructs was compared to eGFP expression from

    mice injected with non-targeting control constructs. In liver, administration of the positive control construct (H1-shRNA(eGFP))

    resulted in a 45% decrease in eGFP transcript levels. Similarly, in liver, constructs AC resulted in 43%, 40% and 54% reductions

    in eGFP transcript levels, respectively. Notably, eGFP suppression with constructs A C did not differ significantly from that

    achieved with the positive control (p = 0.57). All four targeting constructs resulted in significant eGFP suppression compared to

    the non-targeting shRNA negative control (p < 0.05). Due to a lack of plasmid delivery outside of the liver, eGFP expression in

    kidney and spleen with all constructs did not differ significantly from the negative controls. A total of eight mice were injected per

    construct. Error bars represent standard deviation

    EIF2AK2) was compared to their expression levels in

    mice transfected with either the H1-shRNA(eGFP) positive

    control (directed towards the same eGFP target) or

    the non-targeting shRNA control. Transcript levels were

    measured by real-time rtPCR and normalized using

    GAPDH, actin and cyclophilin-A expression. For the

    interferon response the 2CT method was used to

    determine mean fold changes in gene expression [13].

    As shown in Figure 8, the resulting data indicated that

    there was no significant change in expression of IFITM1

    (p = 0.33) and EIF2AK2 (p = 0.11) with constructs AC

    when compared to the non-targeting control construct.Although OAS1 expression varied between 2.3- to 3.0-fold

    compared to the non-targeting control construct, there

    was no significant (p = 0.49) change in OAS1 expression

    compared to the H1-shRNA(eGFP) positive control.

    Furthermore, ISGF3g expression increased between 1.3-

    and 2.0-fold with constructs AC; however, this increase

    in ISGF3g expression was not significant (p = 0.87)

    compared to the H1-shRNA(eGFP) positive control.

    Notably, these results are similar to the expression

    patterns found in vitro. In conclusion, the results

    obtained suggest that, despite the inclusion of additional

    components in constructs AC such as cis-actinghammerhead ribozymes, these constructs operate in a

    similar manner to standard H1-driven shRNA constructs

    in terms of the immune response they elicit in vivo.

    As discussed above, in contrast to the suppression

    obtained in liver, no significant eGFP suppression was

    obtained in kidney or spleen, after tail-vein injection,

    irrespective of the construct used (Figure 6). Delivery

    of naked DNA has been explored extensively for many

    tissues but typically DNA is subject to rapid degradation

    by nucleases [15]. However, results from previous studies

    suggest that hydrodynamic tail-vein injection in mice may

    be used to deliver plasmid DNA effectively to liver [16].

    Absence of eGFP suppression in organs other than liver in

    the current study may be due to inadequate delivery of

    plasmid DNA to those organs. To explore this hypothesisa CMV promoter-eGFP reporter gene construct (CMV-

    eGFP) was delivered to wild-type CD-1, C57BL/6 and

    129 mice by hydrodynamic tail-vein injection and liver,

    kidney, spleen, heart and lung tissues analyzed for eGFP

    expression 48 h post-administration of DNA by real-time

    rtPCR and fluorescent microscopy. eGFP expression was

    found to be almost exclusively limited to liver tissue (data

    not shown).

    In summary, while high-pressure tail-vein administra-

    tion of naked DNA can provide effective in vivo delivery to

    mouse liver, delivery to other organs is restricted or absent

    using this approach. A prerequisite for in vivo explo-ration of specific RNAi is a means of delivering tsRNAi

    constructs to other target organs. Methodologies which

    provide delivery of DNA to organs other than liver have

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    9/12

    Polymerase II Promoter-Driven RNAi 295

    Figure 7. The four panels shown in (A) represent 100-m vibratome liver sections (12.5), under UV light and a GFP filter,

    from C57BL/6-Tg(ACTB-EGFP)1Osb/J mice tail vein injected with a total of 40 g of a non-targeting H1-shRNA and non-targeting

    constructs AC (AD, respectively). A total of 64 mice were treated twice over a 24-h period with 20 g of plasmid (total of 40 g).

    A significant level of green fluorescence can be observed in each of the sections. In these sections it is possible to view an entire liver

    cross-section demonstrating the even distribution of eGFP fluorescence. The four panels in (B) represent 100-m vibratome liver

    sections (12.5), under UV light and a GFP filter, from C57BL/6-Tg(ACTB-EGFP)1Osb/J mice tail vein injected with a total of 40g

    of eGPF targeting H1-shRNA and eGFP-targeting constructs AC (AD, respectively). A significant reduction in green fluorescence

    can be observed in each of the sections in (B) compared to those in (A). Again in these sections it is possible to view an entire

    liver cross-section demonstrating the even distribution of eGFP suppression achieved subsequent to hydrodynamic injection of the

    eGPF-targeting design C construct. A total of eight mice were injected per construct. Scale bars: 500 m

    been employed. For example, Wu and colleagues demon-

    strated plasmid-based delivery of a luciferase reporter

    gene to mouse kidney subsequent to a single injection

    into the inferior vena cava (IVC) [17]. Similarly, in the

    current study, a CMV-driven eGFP plasmid (40 g) has

    been injected into mouse IVC and eGFP expression eval-

    uated using real-time rtPCR and fluorescent microscopy.

    eGFP expression was observed at both RNA and protein

    levels (data not shown) confirming that IVC injection

    may be used to deliver plasmid DNA to mouse kid-

    ney.Given that constructs A C incorporating cis-acting

    hammerhead ribozymes and shRNA sequences have been

    found to generate potent siRNAs in vivo from polymerase

    II promoters, together with methods for delivery of

    DNA to multiple organs, it was timely to engineer a

    liver-specific promoter sequence into these constructs.

    The CMV promoter in constructs AC was replaced by

    2.3 kb of the rat albumin promoter/enhancer sequence,

    a promoter previously shown to drive liver-specific gene

    expression in vivo [18]. Subsequently, eGFP-targeting and

    non-targeting construct C driven by an albumin promoter

    were administered by hydrodynamic tail-vein injection. A total of 24 mice were injected. Liver from injected

    mice were harvested 48 h post-injection and analyzed

    for eGFP expression using real-time rtPCR (Figure 9).

    eGFP expression levels were compared to levels in mice

    injected with non-targeting control plasmids. Significant

    suppression (p < 0.05, DataDesk v6.0) of eGFP expression

    was observed approximately 56% suppression of eGFP

    expression as evaluated by real-time rtPCR.

    In addition, the CMV-driven and albumin-driven eGFP-

    targeting and non-targeting construct C plasmids were

    injected into the IVC. Approximately 45% suppression

    (p < 0.05, DataDesk v6.0) of eGFP expression as

    evaluated by real-time rtPCR was observed in mouse

    kidney (Figure 10) when the CMV-driven eGFP-targetingconstruct was injected andcompared to that obtained with

    a non-targeting control. In contrast, the albumin-driven

    construct elicited no suppression of eGFP expression

    in mouse kidney. In summary, the albumin promoter

    in concert with hammerhead ribozymes and shRNA

    sequences can be used to drive expression of functional

    siRNAs in liver.

    Discussion

    In this study a method of generating functional siRNAs, inprinciple from any polymerase II promoter, involving use

    of cis-acting ribozymes has been developed and evaluated

    both in vitro and in vivo (in mice). As demonstrated such

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    10/12

    296 D. Allen et al.

    Figure 8. Graphs of OAS1 (A), IFITM1 (B), ISGF3g (C), and EIF2AK2 (D) transcript levels in liver from C57BL/

    6-Tg(ACTB-EGFP)1Osb/J mice tail vein injected with constructs AC, the H1-shRNA(eGFP) or the non-targeting shRNA control

    and an eGFP reporter plasmid. A total of 40 mice were treated twice over a 24-h period with 20 g of plasmid (total of 40 g).

    Transcript levels were measured by real-time rtPCR and normalized using GAPDH, actin and cyclophilin-A expression. Expression

    was compared to cells transfected with either the H1-shRNA(eGFP) positive control (directed towards the same eGFP target) or the

    non-targeting shRNA control. The resulting data indicated that there was no significant change in expression of IFITM1 ( p = 0.33)

    and EIF2AK2 (p = 0.11) with constructs AC when compared to the non-targeting control construct. Although OAS1 expression

    varied between 2.3- to 3.0-fold compared to the non-targeting control construct, there was no significant (p = 0.49) change in OAS1expression compared to the H1-shRNA(eGFP) positive control. Furthermore, ISGF3g expression increased between 1.3- and 2.0-fold

    with constructs AC; however, this increase in ISGF3g expression was not significant (p = 0.87) compared to the H1-shRNA(eGFP)

    positive control. A total of eight mice were injected per construct. Error bars represent standard deviation

    an approach may readily be adapted to provide tissue-

    specific RNAi by incorporating tissue-specific polymerase

    II promoters into constructs AC. Potential advantages

    of tissue-specific gene silencing are evident given

    results from microarray-based expression profiling studies

    highlighting that off-target effects can frequently arise

    as a result of RNAi-mediated gene silencing [12].

    Additionally, RNAi has been proposed as a means ofgenerating knockout or knockdown transgenic animals

    [19]. Tissue-specific RNAi would overcome potential

    embryonic lethality in such transgenics and moreover

    would provide enhanced resolution for RNAi technology

    enabling gene silencing in individual tissue types. Given

    these significant advantages a number of approaches to

    achieve tissue-specific expression of functional siRNAs

    have been proposed. The approaches typically utilize H1

    or U6 promoter-driven shRNA constructs carrying loxP

    sites in combination with tissue-specific delivery of Cre

    recombinase to induce expression of functional siRNAs

    [20 23]. Furthermore, tetracycline-based systems havebeen used to demonstrate conditional suppression in vitro

    and in vivo [24,25]. In one such approach T7 recombinase

    was used in conjunction with T7 promoter-driven shRNAs

    and a 3 ribozyme to process shRNAs to achieve

    inducible RNAi [25]. In addition, artificially generated

    microRNAs (miRNAs) have recently been explored in vitro

    as an alternative means of achieving gene silencing

    and potentially could be used in combination with a

    variety of polymerase II promoters to provide tissue

    specificity [26,27]. The strategy explored in vitro and

    in vivo in the current study represents an alternativemeans of achieving siRNA-based tissue-specific RNAi

    and, in contrast to many of the approaches referred

    to above, does not require delivery of Cre recombinase

    or doxycycline to elicit tissue-specific expression and

    would simply utilize a single cassette to achieve tissue-

    specific RNAi. In essence, the approach developed

    incorporates the use of polymerase II promoters to drive

    tissue-specific expression in conjunction with cis-acting

    hammerhead ribozymes to trim the sense and antisense

    strands of the dsRNA to produce functional siRNA or

    shRNA. Given the almost ubiquitous application of RNAi

    in many areas of molecular biology, it is clear thatthere will be multiple applications for systems such as

    that described which augment the resolution of RNAi

    technology.

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    11/12

    Polymerase II Promoter-Driven RNAi 297

    Figure 9. Graph of eGFP transcript levels in liver of

    C57BL/6-Tg(ACTB-EGFP)1Osb/J mice. The albumin promoter-

    driven eGFP-targeting and non-targeting construct C were eval-

    uated in vivo subsequent to hydrodynamic tail-vein injection. A

    total of 16 mice were treated twice over a 24-h period with 20 g

    of plasmid (totalof 40 g). eGFP transcript levels were measured

    by real-time rtPCR and normalized using GAPDH and riboso-

    mal 18s expression. eGFP expression in mice injected with the

    eGFP-targeting construct was compared to eGFP expressionfrom

    mice injected with the non-targeting control construct. In liver,

    administration of the albumin promoter-driven eGFP-targeting

    construct resulted in a 56% decrease in eGFP transcript levels.

    Delivery of this construct resulted in significant eGFP suppres-

    sion compared to the non-targeting shRNA negative control

    (p < 0.05). A total of eight mice were injected per construct.

    Error bars represent standard deviation

    Acknowledgements

    We would like to thank Sylvia Mehigan and Caroline Woods

    who assisted with animal work. The research was supported

    by Enterprise Ireland, DEBRA Ireland and the British Retinitis

    Pigmentosa Society.

    References

    1. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC.Potent and specific genetic interference by double-stranded RNAin Caenorhabditis elegans. Nature 1998; 391: 806811.

    2. Elbashir SM, Harborth J, Lendeckel W, Yalcin A, Weber K,Tuschl T. Duplexes of 21-nucleotide RNAs mediate RNA

    interference in cultured mammalian cells. Nature 2001; 411:494498.

    3. Czauderna F, Fechtner M, Dames S. et al. Structural variationsand stabilising modifications of synthetic siRNAs in mammaliancells. Nucleic Acids Res 2003; 31: 27052716.

    4. Brummelkamp TR, Bernards R, Agami R. A system for stableexpression of short interfering RNAs in mammalian cells. Science2002; 296: 550553.

    5. Miyagishi M, Taira K. U6 promoter-driven siRNAs with foururidine 3 overhangs efficiently suppress targeted geneexpression in mammalian cells. Nat Biotechnol 2002; 20:497500.

    6. Paul CP, Good PD, Winer I, Engelke DR. Effective expression ofsmall interfering RNA in human cells. Nat Biotechnol 2002; 20:505508.

    7. Sui G, Soohoo C, Affar el B, et al. A DNA vector-based RNAi

    technology to suppress gene expression in mammalian cells.Proc Natl Acad Sci U S A 2002; 99: 55155520.8. Yu JY, DeRuiter SL, Turner DL. RNA interference by expression

    of short-interfering RNAs and hairpin RNAs in mammalian cells.Proc Natl Acad Sci U S A 2002; 99: 60476052.

    Figure 10. Graph of eGFP transcript levels in kidney of

    C57BL/6-Tg(ACTB-EGFP)1Osb/J mice. Both the CMV and

    albumin promoter-driven eGFP-targeting and non-targeting

    construct C were evaluated in vivo subsequent to direct inferior

    vena cava injection. A total of 32 mice were treated with40 g of plasmid. eGFP transcript levels were measured by

    real-time rtPCR and normalized using GAPDH and ribosomal

    18s expression. eGFP expression in mice injected with the

    eGFP-targeting construct was compared to eGFP expression from

    mice injected with the non-targeting control construct. In kidney,

    administration of the CMV promoter-driven eGFP-targeting

    construct resulted in a 46% decrease in eGFP transcript

    levels. Delivery of this construct resulted in significant

    eGFP suppression compared to the non-targeting shRNA

    negative control (p < 0.05). In contrast, administration of the

    albumin promoter-driven eGFP-targeting construct resulted in

    no significant decrease in eGFP transcript levels. A total of eight

    mice were injected per construct. Error bars represent standard

    deviation

    9. Kawasaki H, Taira K. Short hairpin type of dsRNAs that arecontrolled by tRNA(Val) promoter significantly induce RNAi-mediated gene silencing in the cytoplasm of human cells. Nucleic

    Acids Res 2003; 31: 700707.10. Paddison PJ, Caudy AA, Hannon GJ. Stable suppression of gene

    expression by RNAi in mammalian cells. Proc Natl Acad Sci U S A2002; 99: 14431448.

    11. Xia H, Mao Q, Paulson HL, Davidson BL. siRNA-mediated genesilencing in vitro and in vivo. Nat Biotechnol 2002; 20:10061010.

    12. Jackson AL, Bartz SR, Schelter J, et al. Expression profilingreveals off-target gene regulation by RNAi. Nat Biotechnol 2003;21: 635637.

    13. Livak KJ, Schmittgen TD. Analysis of relative gene expression

    data using real-time quantitative PCR and the 2(-delta deltaC(T)) method. Methods 2001; 25: 402408.

    14. Okabe M, Ikawa M, Kominami K, Nakanishi T, Nishimune Y.Green mice as a source of ubiquitous green cells. FEBS Lett1997; 407: 313319.

    15. Pouton CW, Seymour LW. Key issues in non-viral gene delivery.Adv Drug Deliv Rev 2001; 46: 187203.

    16. Lewis DL, Hagstrom JE, Loomis AG, Wolff JA, Herweijer H.Efficient delivery of siRNA for inhibition of gene expressionin postnatal mice. Nat Genet 2002; 32: 107108.

    17. Wu X, Gao H, Pasupathy S, Tan PH, Ooi LL, Hui KM. Systemicadministration of naked DNA with targeting specificity tomammalian kidneys. Gene Ther 2005; 12: 477486.

    18. Postic C, Shiota M, Niswender KD, et al. Dual roles forglucokinase in glucose homeostasis as determined by liverand pancreatic beta cell-specific gene knock-outs using Cre

    recombinase. J Biol Chem 1999; 274: 305315.19. Tiscornia G, Singer O, Ikawa M, Verma IM. A general method forgene knockdown in mice by using lentiviral vectors expressingsmall interfering RNA. Proc Natl Acad Sci U S A 2003; 100:18441848.

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm

  • 8/6/2019 Allen et al., 2007

    12/12

    298 D. Allen et al.

    20. Fritsch L, Martinez LA, Sekhri R, et al. Conditional gene knock-down by CRE-dependent short interfering RNAs. EMBO Rep2004; 5: 178182.

    21. Tiscornia G, Tergaonkar V, Galimi F, Verma IM. CRErecombinase-inducible RNA interference mediated by lentiviral

    vectors. Proc Natl Acad Sci U S A 2004; 101: 73477351.22. Ventura A, Meissner A, Dillon CP, et al. Cre-lox-regulated

    conditional RNA interference from transgenes. Proc Natl AcadSci U S A 2004; 101: 10380 10385.23. Coumoul X, Shukla V, Li C, Wang RH, Deng CX. Conditional

    knockdown of Fgfr2 in mice using Cre-LoxP induced RNAinterference. Nucleic Acids Res 2005; 33: e102.

    24. Szulc J, Wiznerowicz M, Sauvain MO, Trono D, Aebischer P. A versatile tool for conditional gene expression and knockdown.Nat Methods 2006; 3: 109116.

    25. Hamdorf M, Muckenfuss H, Tschulena U, et al. An inducibleT7 RNA polymerase-dependent plasmid system. Mol Biotechnol2006; 33: 1321.

    26. Dickins RA, Hemann MT Zilfou JT, et al. Probing tumor

    phenotypes using stable and regulated synthetic microRNAprecursors. Nat Genet. 2005; 37(11): 12891295Oct 2.27. Zeng Y, Cai X, Cullen BR. Use of RNA polymerase II to transcribe

    artificial microRNAs. Methods Enzymol 2005; 392: 371380.

    Copyright 2007 John Wiley & Sons, Ltd. J Gene Med 2007; 9: 287298.

    DOI: 10.1002/jgm