Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Jan 1.
Published in final edited form as: Adv Drug Deliv Rev. 2010 Dec 13;63(1-2):47–68. doi: 10.1016/j.addr.2010.11.003

RNA Interference for improving the Outcome of Islet Transplantation

Feng Li 1, Ram I Mahato 1,*
PMCID: PMC3065652  NIHMSID: NIHMS258415  PMID: 21156190

Abstract

Islet transplantation has the potential to cure type 1 diabetes. Despite recent therapeutic success, it is still not common because a large number of transpanted islets get damaged by multiple challenges including instant blood mediated inflammatory reaction, hypoxia/reperfusion injury, inflammatory cytokines, and immune rejection. RNA interference (RNAi) is an novel strategy to selectively degrade target mRNA. The use of RNAi technologies to downregulate the expression of harmful genes has the potential to improve the outcome of islet transplantation. The aim of this review is to gain a thorough understanding of biological obstacles to islet transplantation and discuss how to overcome these barriers using different RNAi technologies. This eventually will help improve islet survival and function post transplantaion. Chemically synthesized small interferring RNA (siRNA), vector based short haripin RNA (shRNA), and their critical design elements (such as sequences, promoters, backbone) are discussed. The application of combinatorial RNAi in islet transplantation is also discussed. Last but not the least, several delivery strategies for enhanced gene silencing are discussed, including chemical modification of siRNA, complex formation, bioconjugation, and viral vectors.

Keywords: RNA interference, Islet transplantation, siRNA, shRNA

I. Introduction

The first attempt to transplant pancreatic tissue was performed 29 years before the clinical introduction of insulin, when a young boy dying of diabetic ketoacidosis had three small pieces of sheep’s pancreas implanted beneath his skin [1]. But the patient died after 3 days. In 1966, the first pancreatic allotransplantation was performed at the University of Minnesota [2]. Since then, more than 30,000 pancreas transplantations were carried out worldwide. In 2008, approximately 1300 patients were transplanted with pancreas in the United States with the majority of them being simultaneous pancreas-kidney transplantations [3]. Although there are some beneficial effects, pancreas transplantation remains an invasive procedure with significant mortality and morbidity [4]. Pancreas transplantation is also limited by the death of organs, and the need for non-specific immune suppression.

Insulin-producing β cells in the islets of Langerhans are the target of autoimmune aggression in type 1 diabetes, and non-β cells and the exocrine pancreas is unaffected. The islets constitute only a tiny fraction (1%) of the whole pancreas and it is functionally unnecessary to transplant the whole pancreas when only its endocrine tissue is required. Islet transplantation provides a less invasive alternative approach for the treatment of type 1 diabetes with reduced antigen load, relative simplicity, and low morbidity. However, since the initiation of clinical islet transplantation from early 1970s, most of these trials were failed, until the breakthrough at the University of Alberta in Edmonton, Canada. The “Edmonton protocol” was published in 2000, reporting that seven patients with type I diabetes became insulin independent after receiving islet transplantation with a prednisone-free protocol [5].

The Edmonton protocol has been replicated and further modified worldwide with many successes. In 2001, the National Center for Research Resources (NCRR) of the National Institutes of Health (NIH) established ten islet resource centers with the mission of (1) providing high quality human islets for both clinical and basic research use; (2) optimizing processes for islet isolation, purification, and storage; and (3) developing the methods for characterization of islets (http://icr.coh.org). In the same year, the Immune Tolerance Network (ITN) with the joint help from the Juvenile Diabetes Research Foundation (JDRF) and NIH began a trial at nine centers worldwide to perform transplants in 40 subjects to determine if the Edmonton results could be reproduced on a broad scale. The ITN aims at developing a core group of excellent clinical islet transplant centers for future testing of tolerance-based trials (http://www.immunetolerance.org). From 1999 to 2008, around 400 patients received allogeneic islet transplantation [6, 7].

Despite these successes, only less than 10% of the recipients remain insulin independent for up to 5 years, and most recipients return to insulin because the islet function decreased over time [8]. After transplantation, Islet grafts face multiple challenges including innate and adaptive immune rejection, toxicity associated with immune suppressive agents, and insufficient islet revascularization [9, 10]. Therefore, islets isolated from two to four donors are needed for single patient islet transplantation [11]. At present, even with expansion in the currently available organ donor pool, islet transplantation will only benefit about 0.5% of potential recipients [12].

Ex vivo genetic modification of islets before transplantation has the potential to overcome several problems associated with islet transplantation [10, 11, 1317]. Growth factor gene expression has the potential to promote islet revascularization; while anti-apoptotic gene expression can reduce inflammation and immune rejection; and prevent islet cell apoptosis. This strategy can help to reduce the number of islets needed for each recipient and prolong the time that a recipient can maintain insulin independence after transplantation. The genetic modification of islets could be the over-expression of protective genes or inhibition of harmful gene expression [18]. Antisense oligonucleotides (ODNs), ribozymes, DNAzymes, and RNA interference (RNAi) are major approaches which are currently being used to sequence specifically reduce or inhibit gene expression. Among these approaches, RNA interference (RNAi) is relatively new and evolutionally conserved biologic process that regulates gene expression using small interfering RNA (siRNA) mediated sequence specific, post-transcriptional gene silencing [19, 20]. Since the discovery of RNAi by Fire and Milo in 1998, it has been widely used as a tool for basic research as well as a potential therapeutics [2124]. The application of RNAi and other approaches have been extensively reviewed [25]. siRNA molecules can recycle between different copies of mRNA, while antisense ODN can only block the translation of one mRNA before its degradation [26]. Therefore, siRNA is more effective than antisense ODN in reducing target gene expression [27]. Miyagishi et al. has demonstrated that the IC50 for siRNA was about 100-fold lower that of the antisense [28]. Similarly, Bertrand et al. has reported that siRNAs are more efficient in silencing target gene and its effect is lasting longer. In addition, RNAi could also be achieved though the delivery of shRNA, which is expressed endogenously and has longer gene silencing effect [27].

In this review, we will discuss the biological obstacles to islet transplantation and potential RNAi targets; principles and design element for RNAi; and delivery approaches.

II. Obstacles to islet transplantation

Since the success of Edmonton protocol, significant progress has been made in human islet transplantation. However, this protocol is still not widely used mainly due to the fact that islets from several donors are needed for a single recipient, since majority of the islets are damaged prior and post transplantation. The destruction of transplanted islets is a complex problem with many facets including enzyme and mechanical damages induced by the isolation process, hypoxia, inflammatory reaction, and immune rejection. Since many of these factors are inter-related, a thorough understanding of these molecular processes will help us to develop more effective therapeutic interventions.

A. Instant blood-mediated inflammatory reaction

Due to non-immunological and innate immunological factors, 50%–70% islets lose their viability immediately after transplantation by instant blood-mediated inflammatory reaction [2931]. This phenomenon is also called “early graft loss” or “primary nonfunction.” The pathophysiology of blood-mediated inflammatory reaction involves platelet binding and activation, coagulation, complement activation, and infiltration of neutrophils, monocytes, and macrophages (Figure 1) [29]. Activation of coagulation could be triggered through both the intrinsic pathway by islet surface collagen residues and the extrinsic pathway by tissue factors or monocyte chemoattractive protein-1 (MCP-1) [32, 33]. Complement is possibly activated through the alternative pathway [34, 35]. Infiltration of these immune cells could be observed within 1 h after coagulation and complement activation [36]. There are multiple interactions between coagulation, complement, and inflammatory cells, which are responsible for the infiltration of inflammatory cells. Neutrophils and monocytes interact with activated platelets through P-selectin over-expressed on platelets [37]. Macrophages could directly be activated by tissue factor, fibrin, or fibrinogen [38]. Soluble C3a and C5a produced through complement activation are potent chemoattractants for neutrophils and macrophages [34]. The effects of blood-mediated inflammatory reaction are multiple. Firstly, both monocytes and macrophages are major phagocytes leading to the destruction of islet cells. Secondly, proinflammatory cytokines lead to islet cell apoptosis and necrosis. Last but not the least, the infiltrating neutrophils and macrophages will eventually induce subsequent adaptive immune response [39].

Figure 1.

Figure 1

Instant blood-mediated inflammatory reaction.

B. Inflammatory cytokine induced islet cell death

Mechanical and enzymatic stresses during the islet isolation, purification and transfection process result in the activation of resident islet macrophages, which release several pro-inflammatory cytokines including interleukin-1β (IL-1β), tumor necrosis factor- α (TNF-α), and interferon-γ (IFN-γ). In addition, infiltrating host immune cells (monocytes, neutrophils) also produce inflammatory cytokines [40]. These proinflammatory cytokines are responsible for the non-specific inflammation and the damage to transplanted islets during the early stage of transplantation. The exact source and timing of proinflammatory cytokine production still remains elusive and greatly depends on the islet isolation, purification and transplantation protocols. For an example, Kupffer cells produce inflammatory cytokines, NO, and reactive oxygen species (ROS), which are toxic to intraportally transplanted Islets [4143]. The release of cytokines by islet resident macrophages is known to occur during in vitro culture prior to transplantation [44]. The levels of IL-1β and TNF-α increase soon after transplantation, and maximum level was reached at 3 h for IL-1β and at 6 h for TNF-α. Depletion of kupffer cells significantly reduced the levels of IL-1β and TNF-α, indicating the role of macrophage in producing inflammatory cytokines [45]. Another study also showed that IL-1β mRNA was detectable in islet immediately after isolation and was increased after transplantation. IL-1β mRNA increased by 9 fold on day 1, by 7 fold on day 3 after transplantation, and decreased to baseline level on day 7 [46]. It has also been reported that the expression of proinflammatory cytokines (including IL-1β and TNF-α) were observed at 8 h after transplantation and declined after 24 h. In the same study, however, IFN-γ was not observed until 48 h post transplantation [47]. Ozasa et al. reported that IFN-γ mRNA was observed in allogeneic transplantation at 1, 3, 5, and 7 days after transplantation with a peak at day 5, however, it was almost absent in syngeneic transplantation[48]. Although the exact source and level of proinflammatory cytokines production varies among different transplantation protocols used, some common intracellular signal pathways are activated. These cytokines activate JNK/p38, NF-κβ, STAT-1, up-regulate the expression of several significant genes, and finally result in β cell death (Figure 2) [3637].

Figure 2.

Figure 2

Extrinsic and intrinsic pathways for islet cell apoptosis.

IL-1β

The binding of IL-1β to its receptor IL-1R1 leads to the docking of the IL-1 receptor accessory protein (IL-1AcP) and followed by the recruitment of IL-1R1 activated kinase (IRAK) through an adaptor protein named myeloid differentiation factor 88 (MyD88). Recruitment of IRAK leads to the activation of TNF-receptor-associated factor-6 (TRAF6), resulting in the phosphorylation and degradation of IκB. NF-κB is then released from inhibitory IκB, translocates into the nucleus, and induces the expression of multiple genes, including IL-1, IL-6, TNF-α, and iNOS [49, 50].

IFN-γ

IFN-γ binds to IFNγ receptor 1 (IFNγR1), which recruits IFNγ receptor 2 (IFNγR2). IFNγR1and IFNγR2 are associated with Janus tyrosine kinase 1 and 2 (JAK1/2) and results in the activation of JAK1 and JAK2. Then signal transducers and activators of transcription 1 (STAT1) is activated by JAK2 and then translocates to the nucleus, where it binds to the regulatory regions of different genes [51]. In addition, the ischemia induced STAT1 also regulates caspase gene expression through the activation of transcription factor IRF-1 [52].

TNF-α

Upon binding of TNF-α, TNF receptors form trimers and undergo conformational change, leading to the exposure of intracellular death domain. The death domain interacts with TNF-α associated death domain (TRADD), which serves as an adaptor to initiate three signal pathways, including the activation of NF-κB, activation of MAPK pathway, and induction of apoptosis [5355]. TNF receptor-associated factor 2 (TRAF2) and receptor-interacting protein (RIP) are associated with TRADD to recruit and activate kinase IKK, which in turn releases NF-κB by degradation of IκB [53]. TRAF2 also activates MAPK pathways and leads to the protein kinases such as c-jun N-terminal kinase (JNK) and p38 [56]. Apoptosis are activated through both MAPK pathway (direct activation of caspase 3) or FADD mediated activation of caspase 8, which in turn leads to the activation of effector caspase, such as caspase 3 [55].

Under in vitro conditions, IL-1β induces functional impairment in the mouse and rat β cells, and prolonged exposure (6–9 days) to IL-1β in combination with IFN-γ and TNF-α leads to human, mouse, and rat β cell death [50]. The fate of β cells after exposure to inflammatory cytokines depends on the type of cytokines, time course and severity of perturbation to the key β cell gene networks [50, 57]. Inflammatory cytokines upregulate or downregulate around 200 genes, which are either protective or deleterious to β cell survival and function [57, 58]. These genes are related to metabolism, signal transduction, and transcription factors, suggesting that β cells are making attempt to adapt to the effect of cytokine exposure [57]. Several of these genes are putative targets of transcription factor NF-γB [59]. Transcriptional activity of NF-γB is related to the expression of pro-apoptotic genes in β cells [60]. The activation of NF-γB is a necessary step for inducible nitric oxide synthase (iNOS) expression and nitric oxide (NO) production [59]. The cytotoxic effect of cytokines is mediated through the production of radical NO as well as other free radicals, such as peroxynitrite and superoxide [46, 61]. Increased iNOS gene expression leads to an early inflammatory process in islet transplantation, suggesting iNOS is an important mediator of graft inflammation and islet damage in the early stage of islet transplantation. Inflammatory cytokines initiate β cell apoptosis mainly though the extrinsic pathway and function through several molecules in the death receptor pathway, such as Fas (CD95), Fas-associated death domain (FADD), and the death-inducing signaling complex (DISC), leading to the activation and release of caspase 8, which in turn activate caspase 3 [6264].

C. Hypoxia and ischemic reperfusion injury

In an intact pancreas more than 15% of the total blood flow goes to islets, which accounts for only 1% of the pancreatic tissue [65]. During isolation and purification process, the capillary network for blood supply is destroyed; therefore, oxygen and nutrients are solely supplied through diffusion. The oxygen tension in portal vein (where the islets are transplanted) is 8–10 mmHg, which is much lower than that in the intact pancreas (PO2 = 40 mmHg) [66, 67]. Although the angiogenesis initiates soon after transplantation, revascularization of islets may take up to 10–24 days [68]. Therefore, newly transplanted islets are under persistent hypoxic conditions prior to and post transplantation, until revascularization process is completed. The reperfusion following prolonged ischemia causes injury, commonly known as ischemia/reperfusion (IR) injury. For an example, IR injury in the liver activates Kupffer cells, which will produce inflammatory cytokines, ROS and NO, resulting in islet cell apoptosis [69]. Hypoxia induces the intrinsic pathway of apoptosis though the disruption of the mitochondrial integrity, induction of cytochrome C release, and the activation of caspase 9 and 3 [6264]. Nutrient deprivation can also induce islet apoptosis through the intrinsic pathway [62].

D. Autoimmune recurrence and allogeneic immune rejection

In addition to hyperacute rejection, acute rejection, and chronic rejection experienced during the transplantation of other organs such as heart, lung, liver, and kidney, transplanted islets are under additional attacks due to autoimmune rejection against β cells in type I diabetic patients. In the following paragraph, we will briefly summarize the immunology of islet transplantation (Figure 3).

Figure 3.

Figure 3

Autoimmune recurrence in type 1 diabetes and activation of T-cells in host immune rejection against transplanted islets. (A) After transplantation, β cell reactive CD8+, CD4+ T-cells, pro-inflammatory cytokines, and other molecules present in the host attack and destroy transplanted islet β cells. (B) Activation of T-cells in host immune response to transplanted islets. Antigen presenting cells (APCs) take and process antigens from donor and present antigens to host T cells to activate host immune response. Dendritic cells, macrophages, passenger leukocytes, and mononuclear cells are major APCs involved in the antigen presentation. Three signals are needed to fully activate T cells: signal 1, recognition of major histocompatibility complex (MHC) and the bound peptide by the T-cell receptor on the host T cells; signal 2, interaction of co-stimulation signals (such as CD40, CD80/86) with their corresponding receptors on host T cells; signal 3, soluble cytokines (IL-12, IL-10, IL-2) further stimulate the proliferation and differentiation of T cells. These interacting cell surface molecules are also RNAi targets for reduced immune rejection.

1. Autoimmune recurrence

Type I diabetes is caused by autoimmune mediated damage of islet β cells. It is characterized by the presence of β cell reactive T-cells, pro-inflammatory cytokines, and other molecules. Transplanted islets are also attacked by the same stresses that destroy host β cells. Among them, CD8+ T cells are one of the major players in destroying β cells. The infiltration of CD8+ T cells is observed in the pancreas of type 1 diabetic patients as well as transplanted pancreas [70, 71]. In addition, activated CD4+ T cells are also involved in the killing of β cells. Reduced insulitis is observed in the absence of CD4+ T cells [72]. The activation of CD8+ T cells is dependent on CD4+ T cells. CD8+ T cells kill target cells through direct contact and mediated by perforin and granzyme B. Perforin can disturb target cell membrane integrity and facilitate the release of granzyme B into the cytoplasm of target cells, whereas granzyme B cleaves pro-apoptotic molecule Bid. Activated Bid sequesters anti-apoptotic Bcl-2 molecules and triggers β cell apoptosis through the intrinsic apoptotic pathway [7375]. Fas is also upregulated in pancreatic β cells of type 1 diabetic patients and causes β cell apoptosis through FasL-Fas interaction [76, 77].

2. Allogeneic immune rejection

Successful transplantation of allogeneic islets is difficult, because transplanted islets are recognized not only by the host immune system, but the adaptive immune rejections are also activated [78]. Failure of allogeneic islet transplantation is correlated with increased T cell reactivity or alloantibody titers [79, 80]. Antigen presenting cells (APCs) process and present peptide fragments of various surface, secreted and shed protein from the donor. APCs from both donor and host (including dendritic cells, macrophages, passenger leukocytes, and mononuclear cells) are involved in the antigen presentation. After maturation, APCs provide three signals to fully activate T cells: signal 1, recognition of major histocompatibility complex (MHC) and the bound peptide by the T-cell receptor on the host T cells; signal 2, interaction of co-stimulation signals (such as CD40, CD80/86) with their corresponding receptors on host T cells; signal 3, soluble cytokines (IL-12, IL-10, IL-2) further stimulate the proliferation and differentiation of T cells [81, 82]. All of these three signals are required; otherwise, it will result in the apoptosis, anergy, and differentiation of T cells. In addition, anti-donor antibodies are also developed in most allograft recipients after immunosuppressive medication is discontinued. The presence of alloantibodies usually lead to the graft failure of islets from donors with the recognized human leukocyte antigen (HLA) allodeterminants [79].

III. Potential targets for gene silencing

As a potent approach for inhibiting aberrant protein expression, RNAi can theoretically be used to silence any gene involved in the biological processes that lead to the damage of islets. However, due to the complexity of pathophysiology of islet transplantation, only some key mediators are potential targets for RNAi-based intervention. A wide variety of potential therapeutic targets are investigated in the past, which are discussed in the following paragraphs (Table 1).

Table 1.

Some potential RNAi targets for improving islet transplantation

Obstacles Solutions RNAi target genes
Immune rejection

Immune tolerant dendritic cells* NF-κB [8487, 117]
IL-12 [87, 88]
CD80 or CD86 [8991]
CD40 [92, 93]
MyD-88 [94, 263]
SOCS[95]

Reduce islet immunogenicity# Human leukocyte antigen [97100]

Ischemia/reperfusion injury & Apoptosis

Block apoptosis cascades# Caspase 3, 8 [102105]
Fas [106109]

Prevent complement activation# C3, C5α receptor [102, 109, 110]

Inhibit inflammatory pathways# iNOS [111114]
TNF-α [109]
NF-κB [116, 117]
*

Targets in the host;

#

targets in the transplanted islets

A. Modulation of immune rejection

Immunosuppressive agents mediated non-specific immunosuppression is commonly used to reduce the immune rejection in islet transplantation, however, even the most islet-friendly immunosuppressive regimen are toxic to islets. For example, the treatment of islets with sirolimus (also known as rapamycin) or tacrolimus (also known as FK-506 or Fujimycin) causes islet apoptosis [83]. Immune modulation with other biological approaches could also provide alternative ways to reduce immune rejection [10].

1. Genetic modification of dendritic cells

Dendritic cells are one of the most important APCs, which play a critical role in the regulation of immune responses. The activation and maturation of T cells relies on the signals from APCs. T cells will undergo apoptosis, if any one of these signals is blocked. DCs are genetically modified to act as a suppressor of immune responses. These DCs are called tolerogenic DC (Tol-DC). Clinically, we could isolate DCs from the host and then administer them back to the host after making them tolerogenic. Inhibition of the genes associated with DC maturation is one of the promising approaches to generate Tol-DC. NF-γB is an essential transcription factor for DC differentiation and maturation. It includes a group of proteins with similar structures. Inhibition of NF-γB pathway has been shown to produce tolerogenic DC [84, 85]. Silencing of RelB, which is a primary NF-γB protein, reduces the expression of MHC-II, CD80, CD86, and finally prevents DC maturation. RelB silenced DCs were able to inhibit antigen-specific alloreactive immune rejection, and decrease antigen specific T cell proliferation [86]. Significantly reduced allograft rejection was achieved after administration of RelB silenced donor DC [86]. Inhibition of the expression of p35, which is a subunit of IL-12, was an effective approach for antigen-specific immune modulation [87, 88]. DCs treated with siRNA against p35 were able to induce potent Th2 deviation of antigen-specific response; reduce alloreactivity by effective inhibition of T-cell proliferation; shift allogeneic T cell polarization from Th1 to Th2. In another study, Tol-DC was generated through the treatment of antisense ODN to inhibit the expression of CD80 or CD86. Allograft survival was prolonged by administration of modified DCs, which could induce T-cell hyporesponsiveness and apoptosis of T-cells [89]. In addition, silencing of CD80 or CD86 with siRNA inhibited T cell responses in mixed lymphocyte reaction (MLR), generated T regulatory cells [90], and improved cardiac allograft survival [91]. Tol-DC could also be generated through the silencing of CD40, which is another critical co-stimulatory molecules [92]. Silencing of CD40 gene expression in DCs resulted in increased IL-4 production, and decreased IL-12 production and allostimulation activity. The Th2 cytokine production from allogenic T-cells was stimulated [93]. MyD88 is a key adaptor of toll-like receptor signaling [83]. It is critical for DC function and proinflammatory cytokine production. The silencing of MyD88 significantly improved the allograft survival [94]. Suppressor of cytokine signaling (SOCS) molecule is an intracellular inhibitor of Janus kinase. Silencing of SOCS gene in DC resulted in reduced allogeneic mixed leukocyte reactions. In addition, systemic administration of SOCS silenced DCs improved the survival of allograft in rat [95].

2. Reducing immunogenicity of donor islets

HLA is one of the primary causes of Th1/HLA class II antigen rejections in solid organ transplantation. Human islet also has an immunogenic potential, which depends on their HLA incompatibility with the recipient. The alloreactivity is correlated with acute rejection and graft failure. In HLA matching human islet transplantation, the alloreactivity is significantly reduced [96]. However, because of the limited sources of islet donor and extensive HLA polymorphisms, the availability of HLA-match donors is extremely limited. In addition, the fact that multiple donors being required to obtain a sufficient number of islets for transplantation makes the donor selection based on HLA more difficult. Therefore, HLA-unmatched islets are used to meet the clinical needs. These transplanted islets undergo aggressive immune rejections. Silencing of HLA expression provides a promising method to reduce the immunogenicity of HLA-unmatched donor islets [97]. Enhanced resistance to allo-reactive T cell toxicity was observed in human cells treated with lentivirus expressing shRNA against pan class I and allele-specific HLA [98]. Silencing of HLA in solid organ transplantation was also reported [99, 100].

B. Preventing ischemic reperfusion injury and apoptosis

There are multiple challenges including ischemia/reperfusion, mechanical and enzymatic stresses during islet isolation and preservation processes. These factors can activate inflammation and intracellular stress signal pathways, and lead to the overexpression of pro-inflammatory mediators, such as proinflammatory cytokines, ROS, caspases, and transcription factors. The consequence is the loss of islet function and viability. In addition, further adaptive immune response against transplanted islets will also be triggered [101].

1. Block the apoptosis

During the isolation and preservation processes, islets are challenged with various stresses. After transplantation, there are multiple factors that could induce the islet cell death. The survival of transplanted islets could be improved by preventing islet β cells apoptosis [62]. Caspases are the most important mediators of apoptosis, which include initiator caspases (caspase 8 and 9), and effector caspases (caspase 3, 7). Inhibition of caspase activity with caspase inhibitors, or anti-apoptotic genes, has been reported to prevent apoptosis of islets and improve their function [62]. The silencing of caspase 3 and/or caspase 8 was able to reduce renal and liver ischemia/reperfusion injury, which is a significant problem in organ transplantation [102104]. Silencing of caspase 8, which is a mediator of extrinsic apoptosis pathway, could minimize Fas or proinflammatory cytokines induced cell apoptosis. Silencing of caspase 3 (which is the converging point for both extrinsic and intrinsic apoptotic pathways) prevented β cell apoptosis triggered by both extrinsic and intrinsic stimulations. We have demonstrated that the silencing of caspase 3 protects islets from inflammatory cytokine induced apoptosis [105]. Delivery of caspase 3 siRNA to rat insulinoma (INS-1E) cells significantly reduced the number of apoptotic cells. Ex vivo transduction of human islets with adenoviral shRNA prior to transplantation reduced caspase 3 activity in islets and significantly improved islet function. All of the mice transplanted with adv-caspase 3 shRNA transduced islets achieved normoglycemia one day posttransplantation and maintained up to 17 days. In contrast, for mice transplanted with untreated islets, where normoglycemia was achieved in only 60% at day 1, and 80% at day 4 post transplantation. Fas: Fas/FasL interaction is another important mechanism for triggering β cell apoptosis [76]. Treating β cells with siRNA-Fas significantly reduced caspase 3 activity and inhibited Fas-mediated β cell apoptosis [106]. In other studies, Fas siRNA has been successfully used to reduce apoptosis in the liver [107, 108] and heart transplantation [109]. Burkhardt et al. used siRNA to silence Fas expression in murine insulinoma cells. Their results showed that siRNA against Fas inhibited cytokine induced Fas mRNA expression and reduced cell surface Fas protein. However, due to the slow turn-over of Fas protein, a complete inhibition was not observed until prolonged incubation with siRNA. In addition, Fas siRNA effectively reduced cytokine induced β cell apoptosis as determined by caspase 3 activity and TUNEL assay [106].

2. Prevent complement activation

Downregulation of complement 3 or complement C5a receptor with siRNA has the potential to prevent renal ischemia-reperfusion injury [102, 110]. In another study, UW solution containing siRNA against complement 3 also showed the ability to protect donor organs in heart transplantation [109]. For islet transplantation, the complement pathway is activated during the ischemia-reperfusion injury and blood-mediated inflammatory reaction. siRNA targeting C3 and C5a receptors might be a potential target for improving the viability of transplanted islets.

3. Inhibit inflammatory pathway

As discussed in the previous section, proinflammatory cytokines activate several inflammatory signal pathways, upregulate multiple key genes, and finally results in islet cell death. The major mediators in this process are potential targets for preventing islet cell death. iNOS gene: iNOS gene expression on islets increased significantly after transplantation. The maximal iNOS gene expression was observed one day after islet transplantation and then declined progressively [46]. Increased iNOS gene expression led to an early inflammatory process in transplanted islets. Nearly, 50% of the cytokine-induced genes are NO dependent, emphasizing the role of this radical in the late effects of cytokines on insulin producing β cells [57]. Therefore, silencing of iNOS gene might reduce iNOS mediated inflammatory response and prevent islet cell death. McCabe et al. used lentiviral vector expressing shRNA to suppress IL-1β-mediated induction of iNOS expression, resulting in significant protection against the cytotoxic effects of IL-1β exposure [111, 112]. We designed siRNA to inhibit rat and human iNOS gene expression [113, 114]. Due to the difference between human and rat iNOS mRNA sequences, different iNOS siRNA were designed for efficient gene silencing. A dose and sequence dependent inhibition of iNOS gene expression and NO production was observed in rat β cell line (INS-1E cell) and human islets. iNOS gene silencing protected β cells from inflammatory cytokine-induced apoptosis and preserved their insulin secretion ability. However, the effect of iNOS gene silencing on the apoptosis of islet was only moderate, as evidenced by 25–30% reduction in caspase 3 activity and in the percentage of apoptotic cells. Since an islet is a cluster of 200–1000 cells, the transfection efficiency of lipid/siRNA complexes on human islets was only 21–28%, in contrast to the high transfection efficiency (>90%) in β cell lines. Therefore, to achieve satisfactory protective effects on human islets, we need to improve the transfection efficiency, which could be achieved by the rationale design of vectors [105, 115] or modifying the delivery approaches. This will be discussed in the later sections.

NF-γB

Transcription factor NF-γB plays an important role in the inflammation and ischemia-reperfusion injury. For example, several genes are putative targets of NF-γB and their expression levels are changed after cytokine exposure [59]. Transcriptional activity of NF-γB is related to the expression of pro-apoptotic genes in cytokine stimulated β cells [60]. Recombinant adeno-associated virus (AAV) encoding shRNA against NF-γB RelA (p65) subunit was evaluated to reduce NF-γB mediated inflammation [116]. The expressed shRNA against RelA significantly reduced p65 protein expression and suppressed IL-8 secretion in a cellular TNF-α induced inflammation model. In another study, siRNA against NF-γB RelB significantly attenuated ischemia-reperfusion injury induced renal dysfunction and protected mice against lethal kidney ischemia [117].

IV. Method of gene silencing

A. Types of RNAi methods

The applications of RNAi generally utilize two types of molecules: chemically synthesized short interfering RNA (siRNA) or vector based short hairpin RNA (shRNA). Although siRNA and shRNA can be applied to achieve similar functional outcomes, they are intrinsically different molecules as discussed below.

1. Chemically synthesized siRNA

siRNA of 19–23 bp can be synthesized chemically in vitro. Advances in solid phase synthesis technique make it possible to produce siRNA with high purity and precise-controlled sequence. Both sense and antisense strands are synthesized separately and annealed to form double stranded siRNA duplex. After being delivered into cytoplasm, siRNA is directly incorporated into RNA-induced Silencing Complex (RISC) to have its function. Several rate-limiting steps for vector based shRNA are avoided in siRNA mediated gene silencing. Therefore, the gene silencing effect of siRNA is potent and can appear in a short time after transfection. This feature also makes the design of siRNA relatively simple, since the factors regarding short hairpin expression and process are not involved. However, the 2’-OH group in siRNA molecules makes them extremely unstable and susceptible to enzymatic degradation. Only transient gene silencing can be achieved with siRNA and it disappears within several days post transfection.

2. Vector-based shRNA

shRNA is an RNA sequence that makes a tight hairpin turn that can be used to silence gene expression via RNAi. Vector-based shRNA utilizes endogenous cellular machinery to function. After entering the cell nucleus, vector-based shRNA with stem loop structure is expressed to produce pri-shRNA in the nuclei, which is then processed into pre-shRNA by an enzyme named Drosha. Pre-shRNA is exported to the cytoplasm by exportin-5, where it is further processed by Dicer (an RNAse III enzyme) to produce functional siRNA [118]. siRNA is then incorporated into RISC. Activated RISC with antisense strand is formed by cleavage of sense strand. Activated RISC recognizes target mRNA with a complementary sequence, and result in the cleavage of target mRNA (perfect complementarities) or translation inhibition (not prefect complementarities) [22].

B. Design elements of siRNA and shRNA

1. siRNA and shRNA sequences

siRNA sequence is the most important factor that determines the level and duration of gene silencing. Because of high sequence specificity, gene silencing efficiency changes dramatically with minor alteration in siRNA sequence. Mutation in the antisense strand usually makes the activated RISC unable to recognize target mRNA. There are several web-based tools available for predicting effective siRNA sequences and empirical rules are incorporated in the siRNA design software to optimize siRNA sequence [119122]. These rules includes: thermodynamic property [123], length of siRNA sequence [124], GC contents, RNA secondary structure, sequence region, single nucleotide polymorphism sites, avoiding repeats and low complex sequence, avoiding off-target effects on other genes or sequences [125, 126]. siRNA sequences predicted by software, however, are not always very effective in silencing a target gene, and sometime the most potent siRNA sequence is missed. Practically, a potent siRNA can be selected through testing of several candidate siRNA sequences designed with bioinformatics tools.

Since shRNA works through the production of siRNA within the cells, there are plenty of common rules applicable to both siRNA and shRNA sequence design. Actually, people have used the siRNA designing software to predict shRNA sequence with success. Due to the similarity between siRNAs algorithm and that of shRNAs, it is not surprising to keep the gene silencing efficiency after converting siRNAs into shRNAs. However, some features are not required for siRNA but are essential for shRNA expression. For an example, strong GC preference at position 11 and the preference for AU but not GC at position 9 might be the different properties for shRNA sequences [123]. The incompatibility of siRNA design algorithm for shRNA design has also been reported [127]. It would be a good idea to design shRNA sequences using shRNA design algorithm. Similar to siRNA design, several shRNA sequences should be experimentally screened to find potent shRNA sequences. An obvious disadvantage of this approach is that the cloning of shRNA into plasmid is a time consuming process. Fortunately, pre-designed shRNA plasmids for most genes are now commercially available. In most cases, potent shRNA sequences for a particular gene could be found by screening several pre-designed shRNAs.

2. Expression cassette for shRNA

Although shRNA sequence design is critical for efficient gene silencing, it is equally important to have a proper shRNA expression cassette, which determines the magnitude, duration and specificity of shRNA expression. In addition, the structure of shRNA backbone also plays an important role in shRNA transcription, processing, and nuclear export. The effect of different promoters and shRNA structures has been extensively investigated to improve the level and duration of gene silencing.

a. Promoters

Promoters refer to DNA sequences that provide a binding site for RNA polymerase and transcription factors. DNA transcription starts after transcription factor binding to promoter sequences and recruitment of polymerase. The level and duration of shRNA expression depends on the properties of promoters used. Three types of promoters, such as constitutive promoter, inducible promoters, and islet specific promoter will be discussed here.

Constitutive promoters

There are three types of constitutive promoters used for shRNA expression, including polymerase III, polymerase II, and polymerase I promoters. U6 and H1 are the two most commonly used polymerase III (Pol III) promoters for shRNA expression. This is because Pol III promoters transcribe small, highly structured RNAs without poly (A) and 5’-cap in mammalian cells [128]. However, Pol II promoters such as CMV [129, 130], U1 promoter [131], and E1b promoter [132] are also used for shRNA expression. Particularly, CMV promoter can be used for co-expression of a cDNA and a shRNA with a single promoter or to drive the expression of a shRNA embedded in a miRNA structure [130]. Usually, a poly (A) signal is needed for the termination of CMV promoter expression. A modified CMV promoter can also be terminated by poly(T) termination signal [129]. To increase the activity of U6 or H1 promoter, hybrid promoters composed of CMV enhancer and U6 or H1 promoter were constructed for shRNA expression [133135]. Polymerase I promoter has also been reported for shRNA expression [136], however, the use of pol I promoter is still not well investigated.

Inducible promoter

High level constitutive expression of shRNA may saturate cellular machinery and cause cell toxicity. In addition, it is undesirable to constitutively silence genes that are essential for cell viability. Therefore, inducible promoters that can silence genes in a tunable pattern were developed. Different conditional RNAi strategies have been reviewed elsewhere [137]. Here, we list some examples to discuss the concept of using inducible promoters for gene silencing. Czauderna and coworkers [138] constructed tetracycline-inducible U6 or 7SK promoters by introducing a tetracycline operator (tet O) sequence between TATA box and transcription start site. shRNA expression was induced by adding tetracycline which could dissociate the repressors from the tet O. Similarly, tetracycline inducible H1 promoter was also constructed to silence β-catenin [139]. An IPTG-H1 promoter has been developed by placing a lac operator between TATA box and transcription start site [140]. Reversible gene silencing was achieved with above expression systems and the gene silencing effect will disappear within 3–4 day after removing the inducers. An alternative way for inducible gene silencing is Cre-LoxP recombination system, which usually results in an irreversible gene silencing. In this system, a Lox-flanked stuffer sequence with termination sequence (five thymidines) is inserted between H1 promoter and transcription start site. After expression of Cre protein, the stuffer sequence is removed due to the recombination of two loxP sites and thus activate the promoter [141]. Hypoxia inducible promoter may have additional advantages for its application in islets, which are usually under the hypoxic condition after isolation. The use of hypoxia inducible promoter could restrict the gene or shRNA expression within hypoxic islets. Therefore, it could avoid unwanted gene expression and silencing on normoxic tissue or organs and prevent potential toxic side effects. Lee et al. constructed pRTP801-Luc or pRTP801-VEGF plasmids bearing a hypoxia inducible promoter. After transfection into rat islets, transgene expression was higher in hypoxic islets than those in normoxic condition [142, 143]. Although there is still little report about the use of hypoxia inducible expression cassette for gene silencing in islets, we believe that its application in islet gene silencing will increase in the future.

Islet-specific promoter

Several islet-specific promoters have been tested to drive the specific transgene expression in islets. Among them, rat insulin promoters are being extensively investigated [106, 144, 145]. To improve the performance of rat insulin 2 gene promoter in isolated pancreatic islets, an expression cassette including 1.5 kb of the porcine INS 5’ UTR and the 3’ UTR of the bovine growth hormone gene was constructed, which showed robust and specific gene expression on islet β cells [146]. Chai et al. have compared a series of rat insulin promoter sequences with various lengths. The highest promoter activities was observed in a modified RIP3.1 promoter including the non-coding regions in exon1, intron1, and part of exon2 of the rat insulin gene 1 as well as the insulin gene promoter region. This promoter showed 5-fold higher activity in INS-1E cells than a full-length RIP promoter or a CMV promoter. In addition, RIP3.1 promoter demonstrated β cell specific gene expression and glucose responsive gene expression [144]. Recently, the promoter for pri-miR-375 has been indentified, which could selectively express pri-miR-375 in pancreatic islets [147]. Since miRNA based shRNAs have been successfully used for shRNA expression, the finding of this promoter provided an option for islet specific shRNA expression.

b. shRNA backbone structure

High level of shRNA expression does not necessarily lead to efficient gene silencing, because nuclear export, dicer enzyme processing and other rate limiting steps are also critical for efficient gene silencing. These processes are greatly influenced by the shRNA structure, which includes the stem length, loop sequence, the sequence of flanking regions. A typical shRNA is composed of a 19–21bp sense sequence and a 19–21bp antisense sequence linked with 6–9bp loop sequence. A termination sequence of 5–6 thymidines is usually included at the end of stem-loop structure, except for shRNA driven by a pol II promoter. Various shRNA structures have been studied to improve gene silencing efficiency. Siolas et al. found that synthetic 29-mer shRNAs were more potent than 19-mer shRNAs, while the loops sequence did not have any effect [148]. In contrast, Jeanson-Leh et al. reported that increase in the stem length from 19-mer to 29-mer did not show any improvement in gene silencing [149]. It was also reported [123] that gene silencing of shRNAs with 19-mer stem was better than those of shRNA with 29-mer stem in the context of a 9-bp loop.

An emerging trend in shRNA design is to incorporate siRNA into a miRNA precursor backbone so that it can be processed efficiently by intracellular miRNA machinery. miR30-based shRNA has been reported to be more efficient than conventional shRNA [150152]. Boden et al. designed a shRNA against HIV-1 transactivator protein TAT with a miR-30 backbone, which showed 80% more potency in silencing target gene than a conventional shRNA [150]. In another study, Li et al. compared 14 pairs of conventional shRNA and mir30-based shRNA against luciferase gene, and 10 pairs of conventional shRNA and mir30-based shRNA against the mouse tyrosine gene. Results indicated that in 11 out of 14 pairs of shRNA against luciferase and all of those for mouse tyrosine gene, a better gene silencing was observed in conventional shRNA rather than in miR30-based shRNA [123]. Boudreau et al. have also demonstrated that optimized shRNA are more efficient than miR30-based shRNA in silencing of several genes, including green fluorescent protein (GFP), spinocerebellar ataxias, and Huntington disease [153]. In our own studies, we also found that miR30-based shRNA against human iNOS genes was less potent compared with conventional shRNA [115]. These discrepancies indicate that current understanding of shRNA design is still limited and further exploration is necessary to develop a satisfactory rule for shRNA design. The use of miRNA backbone for shRNA expression also has the potential to reduce toxicity [154] and to have tissue or cell type specific gene silencing [155].In addition, we could also construct pri-miRNA cluster, in which multiple miRNA based shRNA is driven by a single promoter to form a single transcript [156, 157].

Recently, several other shRNA structures have been investigated for different applications. In one study, a tRNA-shRNAs chimeric expression cassette was constructed, in which a tRNA promoter was used to drive the expression of shRNA linked to the 3’ acceptor stem of tRNA [158]. This system could achieve similar or even high gene silencing than shRNAs expressed from U6 or H1 promoters [158160]. Since exportin-t, but not exportin-5, was used for the nuclear export of tRNA-shRNAs, it could reduce toxicity related to exprortin-5 saturation [159, 161]. Tandem siRNAs: Zheng et al. designed a dual promoter system (pDual) by inserting siRNA sequence between two promoters (including mouse U6 and human H1 promoters) with opposing directions [162].

3. Off-target effects

Since the first report by Jackson et al. in 2003 [163], many studies have shown that both siRNA and shRNA could cause off-target effects. From a therapeutic standpoint, the off-target effects should be avoided to prevent unintended effects. Therefore, understanding the mechanisms of off-target effects and developing strategies to enhance RNAi specificity is critical for the application of RNAi. There are two types of off-target effects: specific off-target effects, which are caused by the partial sequence complementarity between mRNA and siRNA, and non-specific off-target effects, which are the outcome of overall translation inhibition due to the activation of immune system and cellular toxicity. The mechanism of off-target effects was reviewed extensively elsewhere [164, 165].

Specific off-target effects

Silencing of non-target genes is possible if there is as little as 8-nt homology at the seed region sequence [166]. The off-target effects may also be observed when mRNA shared 7-nt homology with the guide siRNA sequence [167]. However, not all the mRNA with this level of homology will be silenced and it depends on the location of complementary region within siRNA and the mRNA. Seed region is the nucleotides 2–7 at the 5′ end of siRNA. Complementarity between the siRNA seed region and the 3’ UTR of mRNA is an important predictor for off-target effects [166]. Several approaches have been developed to avoid the specific off-target effects. It could be prevented by designing siRNA sequences to minimize the complementarity between seed region of siRNA and the 3’ UTR of mRNA [168]. One or two base mismatch between siRNA and target mRNA are usually tolerated without losing gene silencing efficiency [169]. This allows us to modify the siRNA sequence to minimize the complementarity with 3′UTRs of the untargeted mRNAs. However, the modification of siRNA sequence is associated with the risk of losing gene silencing efficiency. Since the passenger strand is not involved in the recognition with target mRNA, modifications of passenger stand is preferred and more tolerant. For example, breaking a passenger strand into two segments significantly reduced off-target effects [170]. RNA-DNA chimeras [171] have also been developed to avoid off-target effects.

Nonspecific off-target effects

The innate immune systems could be activated by several elements including dsRNA, plasmid or viral vectors, cationic delivery carriers. The activation of immune system causes the inhibition of global gene expression and false positive gene silencing effect. There are multiple receptors involved in the activation of immune system. They are cytoplasmic receptors, including dsRNA-dependent protein kinase R (PKR), retinoic acid-inducible gene-1 (RIG-1), and melanoma differentiation-associated gene-5 (MDA-5) [172174] and endosomal toll-like receptors (TLRs) [175, 176]. For an example, introduction of long dsRNA into cells leads to the activation of PKR, and results in the activation of innate immune system and type-I IFN production [174]. Although the use of shorter siRNAs could minimize the activation of immune response, sequence dependent activation of TLR7 and 8 could not be eliminated [177]. For vector based shRNA, siRNA expressed inside the nucleus and exported into the cytoplasm. The use of vectors based shRNA could thus avoid the activation of TLR3, which recognize dsRNA. However, the plasmid vectors will activate TLR9, unless the unmethylated CpG motifs were removed from the vectors [178, 179]. High levels of siRNA and shRNA could also compete for intracellular miRNA/shRNA process machinery. The saturation of epxortin-5 and RISC component, Argonaute-2 will cause cell toxicity [180]. This can be minimized by designing of proper expression cassettes to carefully control shRNA expression levels and avoid the saturation of miRNA machinery.

V. Combinatorial RNAi

Due to the complexity of molecular mechanism of islet cell death, it is almost impossible to effectively prevent islet cell apoptosis by silencing the expression of a single gene or over-expression of a single protective gene. Therefore, combinatorial RNAi strategies which could synergistically work on multiple targets are required to achieve satisfactory outcomes. This could be realized by simultaneous knockdown of different genes involved in multiple biological processes or signal pathways. Alternatively, we could also simultaneously silence certain harmful genes and at the same time over-express protective genes. Different strategies for combinatorial RNAi will be discussed (Figure 4).

Figure 4.

Figure 4

Combinatorial RNAi strategies for islet genetic modification. (A) siRNA pool targeting different sites of a single gene or targeting multiple genes. (B) Co-expression of multiple shRNAs, (I) multiple shRNAs under separate promoters, (II) miRNA cluster mimics, (III) long hairpin RNA. (C) Co-expression of shRNAs and cDNAs, (I) from a single miRNA backbone, (II) from separate promoters.

A. siRNA Pool

The use of several siRNAs target different regions of a single gene is more efficient than the use of a single siRNA sequence in reducing gene silencing. Cheng et al. tested a pool of two siRNA sequences against TGF-β1 gene in HSC-T6 cells [181]. In that study, the siRNA pool showed better inhibition of TGF- β1 protein expression as determined by western blot analysis. In addition, an enhanced inhibition of the expression of downstream genes including type α1 (I) collagen and alpha-smooth muscle actin (α-SMA) were also observed. In another study, Chen et al. used siRNA pool targeting different regions of human hepatitis B surface antigen. The treatment of this siRNA pool showed efficient inhibition of hepatitis B surface antigens production at HBV-producing HepG2.2.15 cells [182]. Alternatively, co-silencing of multiple genes with siRNAs has also been used to achieve synergistic effect. For an example, siRNA against three different genes including TNF-α, complement 3 (C3) and Fas were tested to reduce ischemia/reperfusion injury to donor organs. The siRNAs used in this study, were supposed to silence several different genes involved in apoptosis, complement activation, and inflammations [109]. In other studies, combination of siRNAs targeting complement 3 and caspase 3 [102] or siRNAs targeting caspase 3 and caspase 8 [103] have been used to prevent renal ischemic injury. siRNA pool targeting mRNAs of multiple genes will be suitable for improving the islet viability because of the involvement of multiple processes in islet damage.

B. Co-expression of multiple shRNAs

There are several scenarios for co-expression of multiple shRNAs from a single vector: (1) multiple shRNAs with the same sequence, (2) multiple shRNAs against different regions of the same gene, or (3) multiple shRNAs against different genes. For the first one, the purpose is to increase the intracellular shRNA levels. However, it is rarely used because the shRNAs levels could be controlled more conveniently by selecting proper promoters. The use of multiple shRNAs against different regions of a single gene is believed to be more efficient in reducing target gene expression levels, possibly due to the synergistic effect. The advantage of using multiple shRNAs against different genes is the possibility to block multiple mediators in the signal pathways. Co-expression of multiple shRNAs were widely studied as an antiviral treatment modality [183]. Although the disease targets are different, the shRNAs expression strategies could be applied to genetic modification of islets. Here, some strategies for co-expression shRNAs will be reviewed.

Gonzalez et al. constructed a vector where multiple shRNAs against HLA class genes were under the control of a common U6 promoter. A dose dependent increase in RNAi efficiency was achieved though raising shRNA copy numbers up to six [184]. Several other studies demonstrated the feasibility of co-expressing multiple shRNAs by utilizing miRNA backbone. In these studies, strong promoters (such as a CMV promoter or a unbiquitin C promoter) were used to control the expression concatemerized multiple miRNA (miR30 or miR155) based shRNAs [156, 185187]. However, the increase in shRNA copy number does not necessary improve gene silencing efficiency. Further investigation will help us to understand the optimal hairpin numbers, positions, backbone sequences required for efficient RNAi.

Long hairpin RNA (lhRNA) [188, 189] or extended hairpin RNAs (e-shRNAs) [190, 191], also showed the potential to co-express several siRNAs. For an example, Liu et al. constructed a vector with e-shRNAs containing two siRNAs and found that a stem length of 43 bp is the minimal requirement. The 66 bp was the minimal length needed to produce three siRNAs [191]. Instead of expressing multiple shRNAs, a single hairpin with multiple siRNAs was expressed in this type of construct. Different levels of gene silencing were observed among these siRNAs and were dependent on their position to the stem loop. The siRNA most close to the loop showed comparable gene silencing to single shRNA, however, other siRNAs usually showed less gene silencing efficiency [189, 191]. In addition, three siRNA units is the upper limit for one construct and the gene silencing efficiency was significantly reduced when four siRNA units were included within one construct[191].

Alternatively, it is also possible to express shRNAs from two or more independent promoters. For example, two shRNAs against different isoforms of glycogen synthase kinase 3 gene were under the control of separate U6 promoters and achieved efficient inhibition of target gene with additive effect [192]. A similar study showed the use of separate U6 promoters to control the expression of shRNAs targeting different Smads [193]. Co-expression of multiple miRNA based shRNAs against different gene has also been reported [194].

C. Co-expression of shRNAs and cDNAs

We have recently demonstrated the beneficial effects of co-expressing multiple genes from a single vector for improving the outcome of islet transplantation. The combination of growth factor and anti-apoptotic genes showed synergistic effect in protecting transplanted islets [16, 17]. A vector could co-express both siRNA and cDNA will be particularly interesting [195]. Not only is the inhibition of harmful genes but also overexpression of protective genes is needed to protect islets. We constructed vectors for co-expression of shRNAs against iNOS genes and a VEGF cDNA. In this vector, VEGF was designed to promote the revascularization of transplanted islets. The shRNA against iNOS were designed to reduce the expression of iNOS genes and minimize iNOS mediated islet cell death. In this study, we also investigated the use different expression cassettes for co-expression of shRNA and cDNA from a single vector. The plasmid vector contained two expression cassettes, where VEGF was under the control of a strong CMV promoter and shRNA was driven by U6, H1, or CMV promoters. Alternatively, it was also possible to express cDNA and shRNA from a single expression cassette. In this case, we used CMV promoter to drive the expression of miRNA based shRNA and VEGF cDNA was inserted between the CMV promoter and shRNA (Figure 5). This strategy was reported previously by Qiu et al., where a miRNA based shRNA was placed in the 3’-UTR of a red fluorescent protein (RFP) for co-expression of both shRNA and RFP [196]. Similar study was also reported by Shin et al., where a GFP gene was inserted between the Tet response element promoter and miR30-based shRNA. This construct allows the co-expression of both GFP and shRNA [194]. In our study, both monocistronic and bicistronic systems showed high VEGF expression and efficient reduction of iNOS gene expression. These vectors could be further modified by replacing VEGF cDNA with another cDNA such as hepatocyte growth factor (HGF), which will enhance angiogenesis and promote β cell proliferation. The shRNA against iNOS genes could be replaced various other shRNA targeting pro-apoptotic and inflammatory genes.

Figure 5.

Figure 5

Bipartite Vectors for co-expression of VEGF cDNA and shRNA against iNOS gene. Silencing of iNOS gene expression with a bipartite vector: (a) with two separate promoters, (b) with shRNA embedded in miRNA backbone and controlled under a single promoter. (c) VEGF expression levels of bipartite vectors. Modified from [115]

VI. Delivery strategies for enhanced gene silencing

Although good progress has been made since the discovery of RNAi technique, the application of RNAi for therapeutic purposes is still limited, because of several biological barriers for siRNA delivery [197]. Systemic delivery of siRNA is required to block host against islet graft immune rejection. However, siRNA face several biological barriers after systemic administration [197], which usually result in the degradation of siRNA. Another potential issue for systemic siRNA delivery is the lack of its targeting ability, which requires the use of targeted delivery system to avoid gene silencing on other tissues or organs. In addition, the systemic administration will possibly induce the immune reaction not only by the siRNA but also by the delivery system, especially, when a viral vector is used. Alternatively, ex vivo delivery is an ideal approach for genetic modification of islets prior to transplantation. Since after isolation, islets are usually preserved in storage solutions before transplantation, we can transfer siRNA into islets prior to transplantation. This ex vivo siRNA delivery approach can also be used to generate immune tolerant dendritic cell (Tol-DC).

The design of proper delivery systems is critical for the clinical application of RNAi technology. Both non-viral and viral vectors have been widely used. The main advantage of viral vectors is their high transduction efficiency. However, the use of viral vectors is undermined due to safety issues and also the construction of individual viral vector is a time consuming process. Alternatively, non-viral approaches, which could avoid the problems for viral vectors, have become a promising approach for gene silencing. In this section, we will discuss siRNA delivery systems applicable to islet transplantation (Figure 6). Firstly, we will discuss the possibility of using free siRNA and chemically modified siRNA. Secondly, we will visit the cationic lipids or polymers for siRNA delivery. Thirdly, the use of bioconjugation will be discussed, which could enhance siRNA delivery into the cells and increase siRNA stability. Finally, we will summarize some viral vectors used for gene silencing including adenovirus and lentivirus (Table II).

Figure 6.

Figure 6

Delivery strategies for gene silencing. Cationic liposomes and polymers are commonly used as transfection reagents for siRNA and shRNA. To avoid the use of cationic carriers, siRNA can also be conjugated to polymers such as poly(ethylene glycol) carrying targeting ligand(s). For enhanced and/or prolonged gene silencing, shRNA is often cloned into a viral (adenovirus, lentivirus, or adeno associate virus) vector. After administration, these delivery systems enter the cells mainly through endocytosis. Vector based shRNA is further translocated into cell nucleus, where shRNA is expressed to produce pri-shRNA, which is further processed into pre-shRNA by an enzyme named Drosha. Pre-shRNA is exported to the cytoplasm by exportin-5, where it is processed by Dicer (an RNAse III enzyme) to produce functional siRNA. For siRNA bioconjugation or siRNA complex, the escape and release of free siRNA from endosome into cytoplasm is an essential step. Finally, siRNAs will be incorporated into RNA-induced silencing complex (RISC) and guide the recognition and degradation of target mRNA.

Table 2.

Delivery strategies for enhance RNAi

Free siRNA

Physical delivery Hydrodynamic injection, In situ perfusion, Microporation

Chemical modification Modifying phosphodiester backbone: Phosphorothioate RNA, boranophosphate RNA, methylphosphonate RNA
Modifying ribose: 2’-O-methyl (2’OMe), 2’-fluoro(2’-F), 2’-O-fluoro-β-D-arabinonucleotide (FANA), 2’-O-(2-methoxyethyl)(MOE), locked nucleic acid

Complex formation

cationic polymers Polyethylenimine, Poly(L-lysine), Chitosan,Cyclodextrin, Poly(disulfide amine), Poly(beta-amino ester)

cationic lipids Lipofectamine-2000, Fugene HD, RNAifect, Solid neleic acid lipid particle (SNALP)

Bioconjugation

siRNA-lipid Cholesterol, Bile acid and long chain fatty acid, Vitamine E

siRNA-peptide Cell penetrating peptides: TAT, penetratin, Transportan

Other conjugates PEG-siRNA, Targeting ligand-PEG-siRNA,Imaging probe(dye)-siRNA

Viral vectors

Adenoviral vectors AdEasy Adenoviral Vector System (Stratagene); Knockout Adenoviral RNAi System (Clontech); BLOCK-iT Adenoviral RNAi Expression System (Invitrogen); AdenoQuick cloning system (OD260)

Lentiviral Vectors BLOCK-iT Lentiviral RNAi Expression System (Invitrogen); Lenti-X shRNA Expression System (clontech); SMARTvector 2.0 Lentiviral shRNA technology (Dharmacon); shRNA libraries TRC lentiviral shRNA library(openbiosysems).

A. Free siRNA

Free siRNAs are difficult to be transfected into the cells mainly because of their poor cell membrane permeability and sensitive to nuclease mediated degradation. However, it is still possible to deliver free siRNAs to islets with some physical approaches, such as hydrodynamic injection [198], in situ perfusion [199], and microporation [200]. Zheng et al. protected donor organs for heart transplantation with siRNA-containing solution [109]. In that study, heart graft from BALB/C mice were preserved in solution containing siRNAs against several genes involved in ischemia/reperfusion injury. This treatment resulted in efficient silencing of target genes in the grafts and lead to improved graft function. Due to the similarity among different organ transplantations, this approach should be equally applicable for islet transplantation.

However, native siRNA is degraded rapidly due to nuclease attack, therefore, chemical modification of siRNA has been extensively studied to (1) minimize nuclease degradation; (2) avoid innate immune system activation; and (3) reduce off-target effects [201, 202]. Since most siRNAs were produced through chemical synthesis, it is technically possible to incorporate various modifications into siRNA backbones (Figure 7). These modifications could be applied to phosphodiester or ribose. Non-bridging oxygen in the phosphodiester could be replaced with sulfur (phosphorothioate), boron (boranophosphate), or methyl (methylphosphonate) groups. The 2’-position of the ribose could also be modified to improve the stability of siRNA. The modification of ribose includes 2’-O-methyl (2’OMe), 2’-fluoro(2’-F), 2’-O-fluoro-β-D-arabinonucleotide (FANA), 2’-O-(2-methoxyethyl) (MOE), and locked nucleic acid (LNA), which contains a methylene bridge connecting the 2’-O with the 4’-C of the ribose ring. Single or combination of various types of modifications could be used depending on the aim of modification, which will be briefly discussed below.

Figure 7.

Figure 7

Common modification of introduced to siRNAs. (A) Modification of non-bridging oxygen internucleotide linkage. (B) Modification of the sugar unit.

Enhancing serum stability

unmodified naked siRNAs are extremely unstable with a serum half-life of less than 5 min [203, 204], which restricts their clinical application. Thus, chemical modification of siRNAs has been extensively studied to improve the resistance of siRNAs to nuclease degradation. Theoretically, all above mentioned modifications can be applied to increase the serum stability of siRNA. The modification strategies should be carefully designed, since extensive modification of siRNAs usually leads to decreased potency [205207]. Alternatively, selective modification of siRNAs at minimal position has been demonstrated to significantly improve the serum stability, while keeping the gene silencing potency. The 2’-O-Me or 2’F-RNA modification have been applied to the termini of the strands or a pyrimidine nucleotide which are vulnerable position for nuclease cleavage [22, 205, 208].

Reducing innate immune system activation

As we discussed previously, siRNAs will activate the innate immune system and result in non-specific off-target effects. Chemical modifications can also be used to avoid or minimize siRNA-mediated activation of immune system. For an example, 2’-O -Me modification of siRNA was able to minimize the immune system activation [209]. Since the 2’-O-Me group is a competitive inhibitor of TLR7, minimal number of 2’-modification is sufficient to inhibit the TLR-7 mediated immune stimulation [210]

Reducing off-target effects

Chemical modification could be applied to reduce specific off-target effects through minimizing the unwanted participation of siRNA in miRNA pathways. For an example, modification of the +2 position with 2’-O-Me on the antisense strand effectively reduced seed region mediated off-target effects [166]. The seed region could also be replaced entirely with DNA residues to reduce off-target effects while preserving the gene silencing potency [171]

B. Complex formation

Both cationic liposomes and polymers are widely used to form complexes with siRNA or plasmid shRNA through electrostatic interaction between positive charged carriers and negatively charged nucleic acids. Complex formation helps to condense siRNA or plasmid DNA and thus facilitate their cellular uptake.

Cationic lipids

Since the introduction of lipofectin at late 1980s [211], many cationic liposomes have been tested for delivery of plasmids, and recently for the delivery of siRNA. There are several commercial cationic lipid formulations for siRNA delivery, including lipofectamine 2000, oligofectamine, lipofectamine, RNAifect, and Fugene HD. We have tested lipofectamine to transfect a plasmid encoding an enhanced green fluorescent protein gene, pCMS-eGFP into intact human islets. However, only low transfection efficiency was observed, this is due to the fact that human islets are a cluster of around 1000 non-dividing cells [14]. In our further studies, a novel lipid with methylsulfonic acid in the cationic head group region to enhance lipid/DNA interaction was synthesized, which showed higher transfection efficiency than those of lipofectamine in intact human islets [212]. In contrast to plasmids, siRNA is efficiently transfected into human islets as well as INS-1E rat β cell line. In this study, we used lipofectamine 2000 to transfect fluorescein-labeled siRNA. The transfection efficiency was determined by flow cytometry analysis (Figure 8). There was increase in transfection efficiency with increase in siRNA concentration, with maximum transfection efficiency of 95.9% and 28.3% on rat β cell line and intact human islets, respectively [113]. This demonstrated the feasibility of in vitro genetically modifying human islets with siRNA/lipid complex.

Figure 8.

Figure 8

Transfection efficiency of siRNA into intact human islets. (A) Fluorescence is seen throughout islet surfaces with some concentrations on edges and interior surface. (B)Following transfection with FITC-labeled siRNA human islets were analyzed by flow cytometry after dispersing into single cells by trypsinization. Compared to the background fluorescence in control group, 21.5 & 28.3% of islet cells incorporated siRNA at 100 & 400 nM concentrations, respectively. Reprinted from [113].

Cationic polymers

In addition to cationic lipids, cationic polymers are also being used for plasmid or siRNA delivery. Although we could achieve decent in vitro transfection efficiency by optimizing transfection conditions, the cellular toxicity of cationic carrier is a major concern for its therapeutic applications. Furthermore, the poor intracellular dissociation of siRNA/polycation complex may reduce the intracellular bioavailability of siRNA. Recently, several biodegradable carriers have been developed to reduce toxicity and improve transfection efficiency. Here, we focus on the some recent progresses, rather than a comprehensive review of the use of cationic polymer for siRNA delivery which was extensively discussed elsewhere [213].

Reducible biodegradable carriers are designed based on the different redox potential between intra and extra cellular environment. The disulfide bonds in these carriers are usually stable enough to keep the integrity of the carrier and protect siRNA before cellular uptake. After cellular uptake, the disulfide bonds are cleaved, in response to the reducing intracellular environment. This will result in the degradation of carriers and release of siRNA. Reducible polymers with disulfide bonds have been investigated for delivery of both plasmid and siRNA [214220]. The advantage of using reducible polymers is obvious, which not only reduced carriers associated toxicity, but also enhanced dissociation of siRNA from polycation/siRNA complex and thus improved gene silencing efficiency. The performance of these carriers could be further improved by incorporation of endosomal escaping moieties, such as histidine [218], and protonatable pendants [219]. Since the endosomal release of siRNA is a significant barrier for siRNA and gene delivery, those components which could facilitate “endosomal escape” will greatly enhance the gene silencing efficiency.

Poly (amino ester) (PAE) is another type of biodegradable polymers which has been extensively studied for gene delivery. Green et al. generated a library of PAE and determined parameters such as polymer type, polymer weight, DNA loading, and other biophysical properties. The leading PAE carrier found in their studies were better than jetPEI and lipofectamine2000 [221]. PAE has also been successfully used for siRNA delivery into lung cancer cells [222], hepatoma cells and primary hepatocytes [223], and fibroblasts [224]. Besides, several other biodegradable polymers can also be used for siRNA delivery, including polyphosphoesters [225227], poly(2-aminoethyl ethylene phosphate) [228], poly(ethylene glycol)-peptide copolymers [229] and ketalized polyethylenimine [230].

C. siRNA bioconjugation

Bioconjugation of siRNA with lipids, polymers or other molecules will help to enhance its cellular uptake, systemic stability, and targeted delivery to specific cells. Technically, siRNA could be modified at the 5’ or 3’ terminus of sense strand or antisense strand. Because of the pivotal role of antisense strand in gene silencing, the modification of sense strand is preferred. Both cleavable and non-cleavable bonds can be used for siRNA conjugation. Acid-sensitive or reducible disulfide bonds are the most commonly used, because of the enhanced release of siRNA from conjugation inside the cell. It could avoid potential negative effects of bioconjugation and provide more amount of free siRNA available for gene silencing.

The molecules used for conjugation with siRNA determine the properties of siRNA conjugate. Attachment of lipids usually increases the hydrophobicity of siRNA, which are highly hydrophilic macromolecules. Lipophilic molecules such as cholesterol [231], bile acid and long chain fatty acid [232, 233], and vitamin E [234] could be conjugated to siRNA. After systemic injection, siRNA-lipid conjugate binds with lipoproteins depending on the degree of hydrophobicity. The interaction between siRNA-lipid conjugate and lipoproteins, lipoprotein receptors and transmembrane proteins has a great influence on the biodistribution and cellular uptake of siRNA after intravenous injection [232]. For an example, siRNA bound to low density lipoprotein (LDL) is mainly delivered to the liver, while a broader distribution was observed for siRNA binding to high density lipoprotein (HDL).

Conjugation of siRNA with cell-penetrating peptides (CPPs) such as TAT [235, 236], penetratin [236] and Transportan [237], has also been studied for siRNA delivery. siRNA-peptide conjugates show a cellular uptake as efficient as that achieved by cationic lipids and resulted in efficient silencing of target genes. However, conjugation of siRNA with peptides does not improve the in vivo stability of siRNA [236]. In addition, innate immune response could be activated after intratracheal administration of pentratin-siRNA conjugate [236], may be due to the immunogenicity of peptide.

To increase the systemic stability of siRNA, poly(ethylene glycol) (PEG) has been conjugated to siRNA by several research groups. In this type of conjugation, siRNA can be linked with PEG through disulfide bond or beta-thio-propionate linkage. Polyelectrolyte complex micelles (PEC micelles) were formed between PEG-siRNA and cationic peptide (KALA) [238] or cationic polymers, such as polyethylenimine (PEI) and poly(L-lysine) (PLL) during the application of siRNA-PEG conjugate. A core was formed between negatively charged siRNA and polycation through condensation, while PEGs are surrounding the core to form a hydrophilic shell. This special core-shell structure can effectively protect siRNA against enzymatic degradation; prevent self-aggregation between PEC micelles, and increase the circulation time after systemic administration. This delivery system can efficiently silence VEGF gene expression in PC-3 cells in vitro [238, 239]. After intravenous injection, enhanced distribution of PEC micelles in tumor was observed, which resulted in significant reduction in VEGF expression in tumor and suppressed tumor growth [240]. siRNA-PEG PEC can be further improved by conjugation of lactose moiety at the distal end of PEG (siRNA-PEG-Lac). This will confer targeting ability to this carrier. The cellular uptake of siRNA in hepatoma cells was enhanced via receptor mediated endocytosis [241, 242]. Delivery of siRNA against RecQL1 gene with lac-PECs inhibited muticellular HuH-7 spheroids growth for up to 21 days, while almost no effect was observed with oligofectAMINE/siRNA lipoplexes. It is probably due to the facilitated cellular uptake of Lac-PECs into the spheroids cells [242]. In addition to the enhanced delivery, siRNA are also conjugated with probes or dyes for bioimaging [243]. Multiple component siRNA conjugate was prepared which includs iron oxide core coated with crosslinked dextran coat, cy5.5 dye, as well as siRNA. After in vitro incubation, this conjugate was efficiently taken up by murine islets. The celluar uptake of siRNA was observed by magnetic imaging of iron core and near-infrared imaging. This study showed the posibility to integrate gene silencing and imaging components into a single conjugate to create a multiple function siRNA conjugtes.

D. Viral vectors

As naturally evolved vehicles, viral vectors utilize the virus infection mechanism for efficient gene transfer into host cells. Although safety concerns retarded the clinical use of viral vectors, their application in genetic modification of islets has the potential to improve the outcome of islet transplantation. This is mainly because of their high transduction efficiency compared with non-viral vectors. Here, we will discuss the properties of some commonly used viral vectors and their application in islet transplantation.

1. Adenovirus

Adenovirus is non-enveloped DNA virus, which could infect both dividing and non-dividing cells. Recombinant adenovirus derived from adenovirus serotype 5 is commonly used for gene delivery. E1 region, which is essential for virus replication, is usually deleted to make replication-deficient adenovirus. In addition, adenoviral genome will not integrate to the targeting cells; therefore, a transient gene expression is achieved by adenoviral vector. The advantages of adenovirus include capacities to carry large DNA insert, easy to produce high titer virus, and its ability to infect non-dividing cells with high efficiency and high gene expression level.

We and others have used replication deficient adenoviral (Adv) vectors for ex vivo transduction of genes to islets to reduce the immune attack, prevent apoptosis, and promote the revascularization of transplanted islets [13, 16, 17, 62, 105, 115, 244]. Bain et al. first reported the feasibility of using Adv vector for gene silencing in islets and β cell lines [245]. Adv vectors expressing siRNA against GLUT2 gene was used in their study to transduce rat islets. Maximum reduction in GLUT2 gene expression was achieved 3 days post transduction with > 90% at both mRNA and protein levels. After then, Adv vectors have been used to silencing different genes in islets to study the function of genes, such as synaptotagmin 9 [246], pyruvate carboxylase [247], tissue factor [248], and proislet amyloid polypeptide [249]. The maximum silencing efficiency was varied from 35% to 75% and achieved 3–4 days post-transduction. Gene silencing efficiency is determined by several factors: (1) the properties (turnover time, abundance) of target gene mRNA and protein; (2) the design of siRNA targeting sequence as well as expression cassette.

Since caspase 3 played an important role in islet β cell apoptosis induced by different factors, the inhibition of caspase activities was a potential approach to improve the graft survival [62]. Recently, we determined the level and duration of caspase-3 gene silencing after transfection of caspase-3 siRNA into INS-1E cells and demonstrated transient gene silencing which did not last beyond 3 days. In an attempt to enhance the level and duration of caspase-3 gene silencing (Figure 9A), we then constructed Adv vectors expressing shRNA against caspase 3 (Adv-caspase-3-shRNA) and determined their gene silencing efficiency in human islets. This adenoviral vector showed efficient caspase-3 gene silencing, which became more significant on day 5 (Figure 9B). Further, Adv-caspase-3-shRNA mediated caspase-3 gene silencing was able to counteract inflammatory cytokine induced islet cell apoptosis in vitro (data not shown). Taking care of these encouraging in vitro results, we then transduced human islets with this advenoviral vector prior to transplantation under the kidney capsules of diabetic NOD-SCID mice. There was some improvement in islet survival and function after transplantation [105].

Figure 9.

Figure 9

Silencing of caspase-3 gene with siRNA and adenoviral shRNA. (a) Effect of caspase-3 siRNA on caspase-3/7 activity in INS-1E cells. After transfection of INS-1E cells with siRNA/Lipofectamine 2000 complexes, the cells were incubated with the cytokine cocktail for additional 16 h. Short term transient caspase 3 were achieved which lasted up to 3 days transfection. (b) Adv-H1-caspase-3-shRNA silencing effect on caspase 3/7 activity in human islets. Upon transduction, islets were incubated in the fresh medium without cytokines, then were collected at the indicated days for determining caspase 3/7 activity. The silencing effect lasted over 5 days post transduction. Modified from [105].

Adv vectors have several particular advantages for gene silencing in human islets. (1) Thorough understanding of the structure of Adv DNA may provide much flexibility in genetically engineer Adv vectors to make them more proper for shRNA delivery and expression. For example, it is convenient to construct shRNA expression vectors with different promoters and backbones with commercially available Adv shuttle vectors for optimal gene silencing; (2) high expression levels achieved by Adv vectors will help to increase gene silencing efficiency and help to get more beneficial effect. (3) Adv vectors has the capacity to carry a large insert, which will make it possible to express multiple shRNAs; or to co-express an shRNA to silence harmful genes and a cDNA to protect islets [115]. The disadvantage of Adv vector is the transient expression profile, which is not suitable for long-term gene silencing. Since the majority of islet cell death occurs within several days after transplantation, the use of Adv to silence genes involved in this process is appropriate.

2. Lentivirus

Lentivirus is a type of retroviral vector, which reverse transcribes into DNA, be actively transported into the nucleus, and integrate into the genome of target cells. Therefore, stable gene expression or silencing could be achieved by using lentiviral vector [250]. Unlike other retrovirus such as moloney murine leukemia virus, which could only transduce dividing cells, lentivirus can transduce both dividing and non-dividing cells. Vesicular stomatitis virus glycoprotein (VSV-G) pseudotyped lentiviral vectors is the most commonly used vector, which could transduce a broad range of cell types [251]. Lymphocytic choriomeningitis virus (LCMV)-pseudotyped is more efficient in transducing human islet insulin-secreting beta cells and shows less toxicity compared with VSV-G pseudotyped lentivirus [252]. Lentivirus were successfully used for genetic modification of islets with various genes, including reporter genes (GFP, Luciferase) [253, 254], antioxidative genes [255], antiapoptotic genes (cFLIP) [253], anti-inflammatory and immunoregulatory genes (CTLA4, TGF-β, interleukin-1 receptor antagonist, and IL-4) [256258]. Unlike Adv vectors which give high level transient gene expression, a prolonged gene expression is usually achieved after transduction islets of lentiviral vectors. For example, after transduction of islets with lentiviral vectors encoding a luciferase gene, bioluminescence is persistent for above 140 days post transplantation. However, the bioluminescence of islets transduced with Adv vector encoding a luciferase gene decrease from thousand fold at the beginning to ten fold over background at 60 days posttransplantation [254]. An optimized ex vivo transduction procedure should neither induce significant toxicity nor affect the insulin release function of islets [255]. In addition, lentiviral vectors induced immune system activation was also minimal [258]. Under optimized conditions, lentivirus could transduce around 10% to 30% of islet cells in an intact islet, which are mainly the cells at the periphery [253, 259, 260]. This is still enough to have significant effect on the viability, probably due to the “barrier” effect, that is shield the core of islets by protecting the periphery of islets [253].

Several other approaches have also been investigated to improve the transduction efficiency. Callewaert et al. demonstrated increased transduction efficiency from around 11.2% to 80.0% by dissociation of islets to single cells prior to transduction and re-aggregation of islets before transplantation [260]. However, improvement in the transfection efficiency did not translate into the improved function of islets. More cells are needed to achieve normoglycemia compared with intact islets. Barbu et al. tested the possibility to transduce islets by whole pancreas perfusion. This transduction procedure could transduce 30% of the cells, while keeping the structural integrity of islets. However, the cells within the core of islets still cannot be effectively transduced [261].

Lentiviral vector has also been used for shRNA delivery for gene silencing [262]. Because of the long term expression achieved by lentiviral vectors, it has been used to stably silence harmful genes to prevent the death of islet β cells. For example, transduction of lentiviral vectors carrying shRNA against iNOS gene, significantly reduced IL-1β induced iNOS expression in islet β cells and protect β cells from IL-1β induced cell death [111]. Since the silencing effect achieved by lentiviral vector is long term, it is suitable for delivery of shRNA against genes that need to be continuously silenced. Otherwise, an inducible or switchable expression cassette should be used to control the timing of gene silencing and thus avoid potential side effects.

VII. Concluding Remarks

Theoretically, any gene of interest could be silenced or effectively inhibited, which is unachievable by conventional small molecule chemicals. In the past few years, RNAi technology has moved rapidly from bench to bedside at unprecedented speed. Currently, there are more than ten ongoing and completed clinical trials using siRNA as therapeutics (http://clinicaltrials.gov/). It is interesting to note that one of these trials sponsored by Quark Pharmaceuticals was related to kidney transplantation, where a siRNA named I5NP was tested for the prophylaxis of delayed graft function in kidney transplantation patients. Although recent progresses have already demonstrated its great potential, the application of RNAi technology for improving the outcome of islet transplantation is still a fledgling area. Further understanding of the biological perspective of the islets transplantation is essential, as it may help to discover effective RNAi target genes. A combination of multiple gene silencing or simultaneously gene silencing and gene over-expression is likely to have improved beneficial effects through synergistic effects. Meanwhile, the advances in RNAi technology will enable us to design more potent siRNA or shRNAs, while minimizing the side effects caused by off-target effects and immune simulation. Last but not the least, the development of safe and efficient delivery systems is the key for successful gene silencing.

Acknowledgments

This research was supported by a grant RO1 DK69968 from the National Institutes of Health (NIH). We thank Dr. Akshay Pratap for critical reading of this manuscript.

List of abbreviations

2’-F

2’-fluoro

2’OMe

2’-O-methyl

α-SMA

alpha-smooth muscle actin

AAV

Recombinant adeno-associated virus

Adv

adenoviral

APCs

Antigen presenting cells

C3

Complement 3

CPPs

Cell-penetrating peptides

DISC

Death-inducing signaling complex

e-shRNAs

Extended hairpin RNAs

FADD

Fas-associated death domain

FANA

2’-O-fluoro-β-D-arabinonucleotide

GFP

Green fluorescent protein

HDL

High density lipoprotein

HGF

Hepatocyte growth factor

HLA

Human leukocyte antigen

IFN-γ

Interferon-γ

IFNγR1

IFNγ receptor 1

IKK

IκB kinase

IL-1AcP

IL-1 receptor accessory protein

IL-1β

Interleukin-1β

iNOS

Inducible nitric oxide synthase

IRAK

IL-1R1 activated kinase

ITN

Immune Tolerance Network

JAK1/2

Janus tyrosine kinase 1 and 2

JDRF

Juvenile Diabetes Research Foundation

JNK

c-jun N-terminal kinase

LCMV

Lymphocytic choriomeningitis virus

LDL

Low density lipoprotein

lhRNA

Long hairpin RNA

LNA

Locked nucleic acid

MCP-1

Monocyte chemoattractive protein-1

MDA-5

Melanoma differentiation-associated gene-5

MHC

Major histocompatibility complex

MLR

Mixed lymphocyte reaction

MOE

2’-O-(2-methoxyethyl)

MyD88

Myeloid differentiation factor 88

NCRR

National Center for Research Resources

NIH

National Institutes of Health

NO

Nitric oxide

PEC micelles

Polyelectrolyte complex micelles

PEG

Poly (ethylene glycol)

PEI

Polyethylenimine

PKR

dsRNA-dependent protein kinase R

PLL

Poly (L-lysine)

RFP

Red fluorescent protein

RIG-1

Retinoic acid-inducible gene-1

RISC

RNA-induced Silencing Complex

ROS

Reactive oxygen species

shRNA

Short hairpin RNA

SOCS

Suppressor of cytokine signaling

STAT1

Signal transducers and activators of transcription 1

tet O

Tetracycline operator

TLR

Toll-like receptor

TNF-α

Tumor necrosis factor-alpha

Tol-DC

Tolerogenic DC

TRADD

TNF-α associated death domain

TRAF2

TNF-receptor-associated factor-2

TRAF6

TNF-receptor-associated factor-6

UTR

Untranslated region

VEGF

Vascular endothelial growth factor

VSV-G

Vesicular stomatitis virus glycoprotein

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Williams PW. Notes on diabetes treated with extract and by grafts of sheep’s pancreas. Br Med J. 1894;2:1303–1304. [Google Scholar]
  • 2.Kelly WD, Lillehei RC, Merkel FK, Idezuki Y, Goetz FC. Allotransplantation of the pancreas and duodenum along with the kidney in diabetic nephropathy. Surgery. 1967;61:827–837. [PubMed] [Google Scholar]
  • 3.Vardanyan M, Parkin E, Gruessner C, Rodriguez Rilo HL. Pancreas vs. islet transplantation: a call on the future. Curr Opin Organ Transplant. 2010;15:124–130. doi: 10.1097/MOT.0b013e32833553f8. [DOI] [PubMed] [Google Scholar]
  • 4.Troppmann C. Complications after pancreas transplantation. Curr Opin Organ Transplant. 2010;15:112–118. doi: 10.1097/MOT.0b013e3283355349. [DOI] [PubMed] [Google Scholar]
  • 5.Shapiro AM, Lakey JR, Ryan EA, Korbutt GS, Toth E, Warnock GL, Kneteman NM, Rajotte RV. Islet transplantation in seven patients with type 1 diabetes mellitus using a glucocorticoid-free immunosuppressive regimen. N Engl J Med. 2000;343:230–238. doi: 10.1056/NEJM200007273430401. [DOI] [PubMed] [Google Scholar]
  • 6.Alejandro R, Barton FB, Hering BJ, Wease S. 2008 Update from the Collaborative Islet Transplant Registry. Transplantation. 2008;86:1783–1788. doi: 10.1097/TP.0b013e3181913f6a. [DOI] [PubMed] [Google Scholar]
  • 7.Merani S, Shapiro AM. Current status of pancreatic islet transplantation. Clin Sci (Lond) 2006;110:611–625. doi: 10.1042/CS20050342. [DOI] [PubMed] [Google Scholar]
  • 8.Ryan EA, Paty BW, Senior PA, Bigam D, Alfadhli E, Kneteman NM, Lakey JR, Shapiro AM. Five-year follow-up after clinical islet transplantation. Diabetes. 2005;54:2060–2069. doi: 10.2337/diabetes.54.7.2060. [DOI] [PubMed] [Google Scholar]
  • 9.Harlan DM, Kenyon NS, Korsgren O, Roep BO. Current advances and travails in islet transplantation. Diabetes. 2009;58:2175–2184. doi: 10.2337/db09-0476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Narang AS, Mahato RI. Biological and biomaterial approaches for improved islet transplantation. Pharmacol Rev. 2006;58:194–243. doi: 10.1124/pr.58.2.6. [DOI] [PubMed] [Google Scholar]
  • 11.Shapiro AM, Ricordi C, Hering BJ, Auchincloss H, Lindblad R, Robertson RP, Secchi A, Brendel MD, Berney T, Brennan DC, Cagliero E, Alejandro R, Ryan EA, DiMercurio B, Morel P, Polonsky KS, Reems JA, Bretzel RG, Bertuzzi F, Froud T, Kandaswamy R, Sutherland DE, Eisenbarth G, Segal M, Preiksaitis J, Korbutt GS, Barton FB, Viviano L, Seyfert-Margolis V, Bluestone J, Lakey JR. International trial of the Edmonton protocol for islet transplantation. N Engl J Med. 2006;355:1318–1330. doi: 10.1056/NEJMoa061267. [DOI] [PubMed] [Google Scholar]
  • 12.Miszta-Lane H, Mirbolooki M, James Shapiro AM, Lakey JR. Stem cell sources for clinical islet transplantation in type 1 diabetes: embryonic and adult stem cells. Med Hypotheses. 2006;67:909–913. doi: 10.1016/j.mehy.2006.03.036. [DOI] [PubMed] [Google Scholar]
  • 13.Cheng K, Fraga D, Zhang C, Kotb M, Gaber AO, Guntaka RV, Mahato RI. Adenovirus-based vascular endothelial growth factor gene delivery to human pancreatic islets. Gene Ther. 2004;11:1105–1116. doi: 10.1038/sj.gt.3302267. [DOI] [PubMed] [Google Scholar]
  • 14.Mahato RI, Henry J, Narang AS, Sabek O, Fraga D, Kotb M, Gaber AO. Cationic lipid and polymer-based gene delivery to human pancreatic islets. Mol Ther. 2003;7:89–100. doi: 10.1016/s1525-0016(02)00031-x. [DOI] [PubMed] [Google Scholar]
  • 15.Narang AS, Cheng K, Henry J, Zhang C, Sabek O, Fraga D, Kotb M, Gaber AO, Mahato RI. Vascular endothelial growth factor gene delivery for revascularization in transplanted human islets. Pharm Res. 2004;21:15–25. doi: 10.1023/b:pham.0000012147.52900.b8. [DOI] [PubMed] [Google Scholar]
  • 16.Panakanti R, Mahato RI. Bipartite vector encoding hVEGF and hIL-1Ra for ex vivo transduction into human islets. Mol Pharm. 2009;6:274–284. doi: 10.1021/mp800183b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Panakanti R, Mahato RI. Bipartite adenoviral vector encoding hHGF and hIL-1Ra for improved human islet transplantation. Pharm Res. 2009;26:587–596. doi: 10.1007/s11095-008-9777-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Mahato RI. Gene expression and silencing for improved islet transplantation. J Control Release. 2009;140:262–267. doi: 10.1016/j.jconrel.2009.04.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998;391:806–811. doi: 10.1038/35888. [DOI] [PubMed] [Google Scholar]
  • 20.Montgomery MK, Xu S, Fire A. RNA as a target of double-stranded RNA-mediated genetic interference in Caenorhabditis elegans. Proc Natl Acad Sci U S A. 1998;95:15502–15507. doi: 10.1073/pnas.95.26.15502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Nguyen T, Menocal EM, Harborth J, Fruehauf JH. RNAi therapeutics: an update on delivery. Curr Opin Mol Ther. 2008;10:158–167. [PubMed] [Google Scholar]
  • 22.de Fougerolles A, Vornlocher HP, Maraganore J, Lieberman J. Interfering with disease: a progress report on siRNA-based therapeutics. Nat Rev Drug Discov. 2007;6:443–453. doi: 10.1038/nrd2310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.De Paula D, Bentley MV, Mahato RI. Hydrophobization and bioconjugation for enhanced siRNA delivery and targeting. Rna. 2007;13:431–456. doi: 10.1261/rna.459807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Chen Y, Cheng G, Mahato RI. RNAi for treating hepatitis B viral infection. Pharm Res. 2008;25:72–86. doi: 10.1007/s11095-007-9504-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Scherer LJ, Rossi JJ. Approaches for the sequence-specific knockdown of mRNA. Nat Biotechnol. 2003;21:1457–1465. doi: 10.1038/nbt915. [DOI] [PubMed] [Google Scholar]
  • 26.Scherer L, Rossi JJ. RNAi applications in mammalian cells. Biotechniques. 2004;36:557–561. doi: 10.2144/04364TE01. [DOI] [PubMed] [Google Scholar]
  • 27.Bertrand JR, Pottier M, Vekris A, Opolon P, Maksimenko A, Malvy C. Comparison of antisense oligonucleotides and siRNAs in cell culture and in vivo. Biochem Biophys Res Commun. 2002;296:1000–1004. doi: 10.1016/s0006-291x(02)02013-2. [DOI] [PubMed] [Google Scholar]
  • 28.Miyagishi M, Hayashi M, Taira K. Comparison of the suppressive effects of antisense oligonucleotides and siRNAs directed against the same targets in mammalian cells. Antisense Nucleic Acid Drug Dev. 2003;13:1–7. doi: 10.1089/108729003764097296. [DOI] [PubMed] [Google Scholar]
  • 29.van der Windt DJ, Bottino R, Casu A, Campanile N, Cooper DK. Rapid loss of intraportally transplanted islets: an overview of pathophysiology and preventive strategies. Xenotransplantation. 2007;14:288–297. doi: 10.1111/j.1399-3089.2007.00419.x. [DOI] [PubMed] [Google Scholar]
  • 30.Bennet W, Groth CG, Larsson R, Nilsson B, Korsgren O. Isolated human islets trigger an instant blood mediated inflammatory reaction: implications for intraportal islet transplantation as a treatment for patients with type 1 diabetes. Ups J Med Sci. 2000;105:125–133. doi: 10.1517/03009734000000059. [DOI] [PubMed] [Google Scholar]
  • 31.Korsgren O, Nilsson B, Berne C, Felldin M, Foss A, Kallen R, Lundgren T, Salmela K, Tibell A, Tufveson G. Current status of clinical islet transplantation. Transplantation. 2005;79:1289–1293. doi: 10.1097/01.tp.0000157273.60147.7c. [DOI] [PubMed] [Google Scholar]
  • 32.Moberg L, Johansson H, Lukinius A, Berne C, Foss A, Kallen R, Ostraat O, Salmela K, Tibell A, Tufveson G, Elgue G, Nilsson Ekdahl K, Korsgren O, Nilsson B. Production of tissue factor by pancreatic islet cells as a trigger of detrimental thrombotic reactions in clinical islet transplantation. Lancet. 2002;360:2039–2045. doi: 10.1016/s0140-6736(02)12020-4. [DOI] [PubMed] [Google Scholar]
  • 33.Johansson H, Lukinius A, Moberg L, Lundgren T, Berne C, Foss A, Felldin M, Kallen R, Salmela K, Tibell A, Tufveson G, Ekdahl KN, Elgue G, Korsgren O, Nilsson B. Tissue factor produced by the endocrine cells of the islets of Langerhans is associated with a negative outcome of clinical islet transplantation. Diabetes. 2005;54:1755–1762. doi: 10.2337/diabetes.54.6.1755. [DOI] [PubMed] [Google Scholar]
  • 34.Bennet W, Sundberg B, Groth CG, Brendel MD, Brandhorst D, Brandhorst H, Bretzel RG, Elgue G, Larsson R, Nilsson B, Korsgren O. Incompatibility between human blood and isolated islets of Langerhans: a finding with implications for clinical intraportal islet transplantation? Diabetes. 1999;48:1907–1914. doi: 10.2337/diabetes.48.10.1907. [DOI] [PubMed] [Google Scholar]
  • 35.Bennet W, Sundberg B, Lundgren T, Tibell A, Groth CG, Richards A, White DJ, Elgue G, Larsson R, Nilsson B, Korsgren O. Damage to porcine islets of Langerhans after exposure to human blood in vitro, or after intraportal transplantation to cynomologus monkeys: protective effects of sCR1 and heparin. Transplantation. 2000;69:711–719. doi: 10.1097/00007890-200003150-00007. [DOI] [PubMed] [Google Scholar]
  • 36.Moberg L, Korsgren O, Nilsson B. Neutrophilic granulocytes are the predominant cell type infiltrating pancreatic islets in contact with ABO-compatible blood. Clin Exp Immunol. 2005;142:125–131. doi: 10.1111/j.1365-2249.2005.02883.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Contreras JL, Eckstein C, Smyth CA, Bilbao G, Vilatoba M, Ringland SE, Young C, Thompson JA, Fernandez JA, Griffin JH, Eckhoff DE. Activated protein C preserves functional islet mass after intraportal transplantation: a novel link between endothelial cell activation, thrombosis, inflammation, and islet cell death. Diabetes. 2004;53:2804–2814. doi: 10.2337/diabetes.53.11.2804. [DOI] [PubMed] [Google Scholar]
  • 38.Szaba FM, Smiley ST. Roles for thrombin and fibrin(ogen) in cytokine/chemokine production and macrophage adhesion in vivo. Blood. 2002;99:1053–1059. doi: 10.1182/blood.v99.3.1053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Kirchhof N, Shibata S, Wijkstrom M, Kulick DM, Salerno CT, Clemmings SM, Heremans Y, Galili U, Sutherland DE, Dalmasso AP, Hering BJ. Reversal of diabetes in non-immunosuppressed rhesus macaques by intraportal porcine islet xenografts precedes acute cellular rejection. Xenotransplantation. 2004;11:396–407. doi: 10.1111/j.1399-3089.2004.00157.x. [DOI] [PubMed] [Google Scholar]
  • 40.Gysemans C, Callewaert H, Overbergh L, Mathieu C. Cytokine signalling in the beta-cell: a dual role for IFNgamma. Biochem Soc Trans. 2008;36:328–333. doi: 10.1042/BST0360328. [DOI] [PubMed] [Google Scholar]
  • 41.Nathan CF. Secretory products of macrophages. J Clin Invest. 1987;79:319–326. doi: 10.1172/JCI112815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Bendtzen K, Mandrup-Poulsen T, Nerup J, Nielsen JH, Dinarello CA, Svenson M. Cytotoxicity of human pI 7 interleukin-1 for pancreatic islets of Langerhans. Science. 1986;232:1545–1547. doi: 10.1126/science.3086977. [DOI] [PubMed] [Google Scholar]
  • 43.Barshes NR, Wyllie S, Goss JA. Inflammation-mediated dysfunction and apoptosis in pancreatic islet transplantation: implications for intrahepatic grafts. J Leukoc Biol. 2005;77:587–597. doi: 10.1189/jlb.1104649. [DOI] [PubMed] [Google Scholar]
  • 44.Arnush M, Heitmeier MR, Scarim AL, Marino MH, Manning PT, Corbett JA. IL-1 produced and released endogenously within human islets inhibits beta cell function. J Clin Invest. 1998;102:516–526. doi: 10.1172/JCI844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Bottino R, Fernandez LA, Ricordi C, Lehmann R, Tsan MF, Oliver R, Inverardi L. Transplantation of allogeneic islets of Langerhans in the rat liver: effects of macrophage depletion on graft survival and microenvironment activation. Diabetes. 1998;47:316–323. doi: 10.2337/diabetes.47.3.316. [DOI] [PubMed] [Google Scholar]
  • 46.Montolio M, Biarnes M, Tellez N, Escoriza J, Soler J, Montanya E. Interleukin-1beta and inducible form of nitric oxide synthase expression in early syngeneic islet transplantation. J Endocrinol. 2007;192:169–177. doi: 10.1677/joe.1.06968. [DOI] [PubMed] [Google Scholar]
  • 47.Gysemans CA, Waer M, Valckx D, Laureys JM, Mihkalsky D, Bouillon R, Mathieu C. Early graft failure of xenogeneic islets in NOD mice is accompanied by high levels of interleukin-1 and low levels of transforming growth factor-beta mRNA in the grafts. Diabetes. 2000;49:1992–1997. doi: 10.2337/diabetes.49.12.1992. [DOI] [PubMed] [Google Scholar]
  • 48.Ozasa T, Newton MR, Dallman MJ, Shimizu S, Gray DW, Morris PJ. Cytokine gene expression in pancreatic islet grafts in the rat. Transplantation. 1997;64:1152–1159. doi: 10.1097/00007890-199710270-00013. [DOI] [PubMed] [Google Scholar]
  • 49.Thomas HE, McKenzie MD, Angstetra E, Campbell PD, Kay TW. Beta cell apoptosis in diabetes. Apoptosis. 2009;14:1389–1404. doi: 10.1007/s10495-009-0339-5. [DOI] [PubMed] [Google Scholar]
  • 50.Eizirik DL, Mandrup-Poulsen T. A choice of death--the signal-transduction of immune-mediated beta-cell apoptosis. Diabetologia. 2001;44:2115–2133. doi: 10.1007/s001250100021. [DOI] [PubMed] [Google Scholar]
  • 51.Tau G, Rothman P. Biologic functions of the IFN-gamma receptors. Allergy. 1999;54:1233–1251. doi: 10.1034/j.1398-9995.1999.00099.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Stephanou A, Brar BK, Scarabelli TM, Jonassen AK, Yellon DM, Marber MS, Knight RA, Latchman DS. Ischemia-induced STAT-1 expression and activation play a critical role in cardiomyocyte apoptosis. J Biol Chem. 2000;275:10002–10008. doi: 10.1074/jbc.275.14.10002. [DOI] [PubMed] [Google Scholar]
  • 53.Chen G, Goeddel DV. TNF-R1 signaling: a beautiful pathway. Science. 2002;296:1634–1635. doi: 10.1126/science.1071924. [DOI] [PubMed] [Google Scholar]
  • 54.Wajant H, Pfizenmaier K, Scheurich P. Tumor necrosis factor signaling. Cell Death Differ. 2003;10:45–65. doi: 10.1038/sj.cdd.4401189. [DOI] [PubMed] [Google Scholar]
  • 55.Rath PC, Aggarwal BB. TNF-induced signaling in apoptosis. J Clin Immunol. 1999;19:350–364. doi: 10.1023/a:1020546615229. [DOI] [PubMed] [Google Scholar]
  • 56.Saklatvala J, Dean J, Finch A. Protein kinase cascades in intracellular signalling by interleukin-I and tumour necrosis factor. Biochem Soc Symp. 1999;64:63–77. [PubMed] [Google Scholar]
  • 57.Kutlu B, Cardozo AK, Darville MI, Kruhoffer M, Magnusson N, Orntoft T, Eizirik DL. Discovery of gene networks regulating cytokine-induced dysfunction and apoptosis in insulin-producing INS-1 cells. Diabetes. 2003;52:2701–2719. doi: 10.2337/diabetes.52.11.2701. [DOI] [PubMed] [Google Scholar]
  • 58.Eizirik DL, Kutlu B, Rasschaert J, Darville M, Cardozo AK. Use of microarray analysis to unveil transcription factor and gene networks contributing to Beta cell dysfunction and apoptosis. Ann N Y Acad Sci. 2003;1005:55–74. doi: 10.1196/annals.1288.007. [DOI] [PubMed] [Google Scholar]
  • 59.Cardozo AK, Heimberg H, Heremans Y, Leeman R, Kutlu B, Kruhoffer M, Orntoft T, Eizirik DL. A comprehensive analysis of cytokine-induced and nuclear factor-kappa B-dependent genes in primary rat pancreatic beta-cells. J Biol Chem. 2001;276:48879–48886. doi: 10.1074/jbc.M108658200. [DOI] [PubMed] [Google Scholar]
  • 60.Ortis F, Cardozo AK, Crispim D, Storling J, Mandrup-Poulsen T, Eizirik DL. Cytokine-induced proapoptotic gene expression in insulin-producing cells is related to rapid, sustained, and nonoscillatory nuclear factor-kappaB activation. Mol Endocrinol. 2006;20:1867–1879. doi: 10.1210/me.2005-0268. [DOI] [PubMed] [Google Scholar]
  • 61.Corbett JA, Sweetland MA, Wang JL, Lancaster JR, Jr, McDaniel ML. Nitric oxide mediates cytokine-induced inhibition of insulin secretion by human islets of Langerhans. Proc Natl Acad Sci U S A. 1993;90:1731–1735. doi: 10.1073/pnas.90.5.1731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Emamaullee JA, Shapiro AM. Interventional strategies to prevent beta-cell apoptosis in islet transplantation. Diabetes. 2006;55:1907–1914. doi: 10.2337/db05-1254. [DOI] [PubMed] [Google Scholar]
  • 63.Hengartner MO. The biochemistry of apoptosis. Nature. 2000;407:770–776. doi: 10.1038/35037710. [DOI] [PubMed] [Google Scholar]
  • 64.Green DR. Apoptotic pathways: paper wraps stone blunts scissors. Cell. 2000;102:1–4. doi: 10.1016/s0092-8674(00)00003-9. [DOI] [PubMed] [Google Scholar]
  • 65.Jansson L. Dissociation between pancreatic islet blood flow and insulin release in the rat. Acta Physiol Scand. 1985;124:223–228. doi: 10.1111/j.1748-1716.1985.tb07655.x. [DOI] [PubMed] [Google Scholar]
  • 66.Sezai S, Sakurabayashi S, Yamamoto Y, Morita T, Hirano M, Oka H. Hepatic arterial and portal venous oxygen content and extraction in liver cirrhosis. Liver. 1993;13:31–35. doi: 10.1111/j.1600-0676.1993.tb00602.x. [DOI] [PubMed] [Google Scholar]
  • 67.Pileggi A, Ricordi C, Alessiani M, Inverardi L. Factors influencing Islet of Langerhans graft function and monitoring. Clin Chim Acta. 2001;310:3–16. doi: 10.1016/s0009-8981(01)00503-4. [DOI] [PubMed] [Google Scholar]
  • 68.Menger MD, Yamauchi J, Vollmar B. Revascularization and microcirculation of freely grafted islets of Langerhans. World J Surg. 2001;25:509–515. doi: 10.1007/s002680020345. [DOI] [PubMed] [Google Scholar]
  • 69.Li X, Chen H, Epstein PN. Metallothionein protects islets from hypoxia and extends islet graft survival by scavenging most kinds of reactive oxygen species. J Biol Chem. 2004;279:765–771. doi: 10.1074/jbc.M307907200. [DOI] [PubMed] [Google Scholar]
  • 70.Sibley RK, Sutherland DE, Goetz F, Michael AF. Recurrent diabetes mellitus in the pancreas iso- and allograft. A light and electron microscopic and immunohistochemical analysis of four cases. Lab Invest. 1985;53:132–144. [PubMed] [Google Scholar]
  • 71.Santamaria P, Nakhleh RE, Sutherland DE, Barbosa JJ. Characterization of T lymphocytes infiltrating human pancreas allograft affected by isletitis and recurrent diabetes. Diabetes. 1992;41:53–61. doi: 10.2337/diab.41.1.53. [DOI] [PubMed] [Google Scholar]
  • 72.Mora C, Wong FS, Chang CH, Flavell RA. Pancreatic infiltration but not diabetes occurs in the relative absence of MHC class II-restricted CD4 T cells: studies using NOD/CIITA-deficient mice. J Immunol. 1999;162:4576–4588. [PubMed] [Google Scholar]
  • 73.Trapani JA, Smyth MJ. Functional significance of the perforin/granzyme cell death pathway. Nat Rev Immunol. 2002;2:735–747. doi: 10.1038/nri911. [DOI] [PubMed] [Google Scholar]
  • 74.Sutton VR, Davis JE, Cancilla M, Johnstone RW, Ruefli AA, Sedelies K, Browne KA, Trapani JA. Initiation of apoptosis by granzyme B requires direct cleavage of bid, but not direct granzyme B-mediated caspase activation. J Exp Med. 2000;192:1403–1414. doi: 10.1084/jem.192.10.1403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Estella E, McKenzie MD, Catterall T, Sutton VR, Bird PI, Trapani JA, Kay TW, Thomas HE. Granzyme B-mediated death of pancreatic beta-cells requires the proapoptotic BH3-only molecule bid. Diabetes. 2006;55:2212–2219. doi: 10.2337/db06-0129. [DOI] [PubMed] [Google Scholar]
  • 76.Moriwaki M, Itoh N, Miyagawa J, Yamamoto K, Imagawa A, Yamagata K, Iwahashi H, Nakajima H, Namba M, Nagata S, Hanafusa T, Matsuzawa Y. Fas and Fas ligand expression in inflamed islets in pancreas sections of patients with recent-onset Type I diabetes mellitus. Diabetologia. 1999;42:1332–1340. doi: 10.1007/s001250051446. [DOI] [PubMed] [Google Scholar]
  • 77.Loweth AC, Williams GT, James RF, Scarpello JH, Morgan NG. Human islets of Langerhans express Fas ligand and undergo apoptosis in response to interleukin-1beta and Fas ligation. Diabetes. 1998;47:727–732. doi: 10.2337/diabetes.47.5.727. [DOI] [PubMed] [Google Scholar]
  • 78.Pericin M, Althage A, Freigang S, Hengartner H, Rolland E, Dupraz P, Thorens B, Aebischer P, Zinkernagel RM. Allogeneic beta-islet cells correct diabetes and resist immune rejection. Proc Natl Acad Sci U S A. 2002;99:8203–8206. doi: 10.1073/pnas.122241299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Campbell PM, Senior PA, Salam A, Labranche K, Bigam DL, Kneteman NM, Imes S, Halpin A, Ryan EA, Shapiro AM. High risk of sensitization after failed islet transplantation. Am J Transplant. 2007;7:2311–2317. doi: 10.1111/j.1600-6143.2007.01923.x. [DOI] [PubMed] [Google Scholar]
  • 80.Olack BJ, Swanson CJ, Flavin KS, Phelan D, Brennan DC, White NH, Lacy PE, Scharp DW, Poindexter N, Mohanakumar T. Sensitization to HLA antigens in islet recipients with failing transplants. Transplant Proc. 1997;29:2268–2269. doi: 10.1016/s0041-1345(97)00327-8. [DOI] [PubMed] [Google Scholar]
  • 81.Morel Y, Truneh A, Sweet RW, Olive D, Costello RT. The TNF superfamily members LIGHT and CD154 (CD40 ligand) costimulate induction of dendritic cell maturation and elicit specific CTL activity. J Immunol. 2001;167:2479–2486. doi: 10.4049/jimmunol.167.5.2479. [DOI] [PubMed] [Google Scholar]
  • 82.Emamaullee J, Toso C, Merani S, Shapiro AM. Costimulatory blockade with belatacept in clinical and experimental transplantation - a review. Expert Opin Biol Ther. 2009;9:789–796. doi: 10.1517/14712590902942284. [DOI] [PubMed] [Google Scholar]
  • 83.Hui H, Khoury N, Zhao X, Balkir L, D'Amico E, Bullotta A, Nguyen ED, Gambotto A, Perfetti R. Adenovirus-mediated XIAP gene transfer reverses the negative effects of immunosuppressive drugs on insulin secretion and cell viability of isolated human islets. Diabetes. 2005;54:424–433. doi: 10.2337/diabetes.54.2.424. [DOI] [PubMed] [Google Scholar]
  • 84.Bonham CA, Peng L, Liang X, Chen Z, Wang L, Ma L, Hackstein H, Robbins PD, Thomson AW, Fung JJ, Qian S, Lu L. Marked prolongation of cardiac allograft survival by dendritic cells genetically engineered with NF-kappa B oligodeoxyribonucleotide decoys and adenoviral vectors encoding CTLA4-Ig. J Immunol. 2002;169:3382–3391. doi: 10.4049/jimmunol.169.6.3382. [DOI] [PubMed] [Google Scholar]
  • 85.Yoshimura S, Bondeson J, Foxwell BM, Brennan FM, Feldmann M. Effective antigen presentation by dendritic cells is NF-kappaB dependent: coordinate regulation of MHC, co-stimulatory molecules and cytokines. Int Immunol. 2001;13:675–683. doi: 10.1093/intimm/13.5.675. [DOI] [PubMed] [Google Scholar]
  • 86.Li M, Zhang X, Zheng X, Lian D, Zhang ZX, Ge W, Yang J, Vladau C, Suzuki M, Chen D, Zhong R, Garcia B, Jevnikar AM, Min WP. Immune modulation and tolerance induction by RelB-silenced dendritic cells through RNA interference. J Immunol. 2007;178:5480–5487. doi: 10.4049/jimmunol.178.9.5480. [DOI] [PubMed] [Google Scholar]
  • 87.Li M, Qian H, Ichim TE, Ge WW, Popov IA, Rycerz K, Neu J, White D, Zhong R, Min WP. Induction of RNA interference in dendritic cells. Immunol Res. 2004;30:215–230. doi: 10.1385/IR:30:2:215. [DOI] [PubMed] [Google Scholar]
  • 88.Hill JA, Ichim TE, Kusznieruk KP, Li M, Huang X, Yan X, Zhong R, Cairns E, Bell DA, Min WP. Immune modulation by silencing IL-12 production in dendritic cells using small interfering RNA. J Immunol. 2003;171:691–696. doi: 10.4049/jimmunol.171.2.691. [DOI] [PubMed] [Google Scholar]
  • 89.Liang X, Lu L, Chen Z, Vickers T, Zhang H, Fung JJ, Qian S. Administration of dendritic cells transduced with antisense oligodeoxyribonucleotides targeting CD80 or CD86 prolongs allograft survival. Transplantation. 2003;76:721–729. doi: 10.1097/01.TP.0000076470.35404.49. [DOI] [PubMed] [Google Scholar]
  • 90.Suzuki M, Zhang X, Zheng X, Vladau C, Li M, Chen D, Min W. Immune modulation through silencing CD80 and CD86 in dendritic cells using siRNA. J Immunol. 2007;178:84. doi: 10.4049/jimmunol.178.9.5480. [DOI] [PubMed] [Google Scholar]
  • 91.Xiang J, Gu X, Qian S, Chen Z. Graded function of CD80 and CD86 in initiation of T-cell immune response and cardiac allograft survival. Transpl Int. 2008;21:163–168. doi: 10.1111/j.1432-2277.2007.00590.x. [DOI] [PubMed] [Google Scholar]
  • 92.Karimi MH, Ebadi P, Pourfathollah AA, Soheili ZS, Samiee S, Ataee Z, Tabei SZ, Moazzeni SM. Immune modulation through RNA interference-mediated silencing of CD40 in dendritic cells. Cell Immunol. 2009;259:74–81. doi: 10.1016/j.cellimm.2009.05.016. [DOI] [PubMed] [Google Scholar]
  • 93.Zheng X, Vladau C, Zhang X, Suzuki M, Ichim TE, Zhang ZX, Li M, Carrier E, Garcia B, Jevnikar AM, Min WP. A novel in vivo siRNA delivery system specifically targeting dendritic cells and silencing CD40 genes for immunomodulation. Blood. 2009;113:2646–2654. doi: 10.1182/blood-2008-04-151191. [DOI] [PubMed] [Google Scholar]
  • 94.Zhu C, Xu H, Zhang G, Lu C, Ji M, Wu W. Myeloid differentiation factor 88-silenced bone marrow-derived dendritic cells exhibit enhanced tolerogenicity in intestinal transplantation in rats. Transplant Proc. 2008;40:1625–1628. doi: 10.1016/j.transproceed.2008.01.070. [DOI] [PubMed] [Google Scholar]
  • 95.Zhu C, Xu H, Zhang J, Wang K, Zhu P. Donor bone-marrow suppressor of cytokine signaling-1-silenced dendritic cells prolong rat intestinal allograft survival. Transplant Proc. 2008;40:3707–3710. doi: 10.1016/j.transproceed.2008.07.131. [DOI] [PubMed] [Google Scholar]
  • 96.Bouwman LH, Ling Z, Duinkerken G, Pipeleers DG, Roep BO. HLA incompatibility and immunogenicity of human pancreatic islet preparations cocultured with blood cells of healthy donors. Hum Immunol. 2005;66:494–500. doi: 10.1016/j.humimm.2005.01.018. [DOI] [PubMed] [Google Scholar]
  • 97.Haga K, Lemp NA, Logg CR, Nagashima J, Faure-Kumar E, Gomez GG, Kruse CA, Mendez R, Stripecke R, Kasahara N, Cicciarelli JC. Permanent, lowered HLA class I expression using lentivirus vectors with shRNA constructs: Averting cytotoxicity by alloreactive T lymphocytes. Transplant Proc. 2006;38:3184–3188. doi: 10.1016/j.transproceed.2006.10.158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Hacke K, Falahati R, Flebbe-Rehwaldt L, Kasahara N, Gaensler KM. Suppression of HLA expression by lentivirus-mediated gene transfer of siRNA cassettes and in vivo chemoselection to enhance hematopoietic stem cell transplantation. Immunol Res. 2009;44:112–126. doi: 10.1007/s12026-008-8088-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Figueiredo C, Seltsam A, Holtkamp U, Blasczyk R. Stable knock down of non-acceptable HLA mismatches in solid organ transplantation. Hum Immunol. 2004;65:S7. [Google Scholar]
  • 100.Figueiredo C, Seltsam A, Blasczyk R. Class-, gene-, and group-specific HLA silencing by lentiviral shRNA delivery. J Mol Med. 2006;84:425–437. doi: 10.1007/s00109-005-0024-2. [DOI] [PubMed] [Google Scholar]
  • 101.Marzorati S, Pileggi A, Ricordi C. Allogeneic islet transplantation. Expert Opin Biol Ther. 2007;7:1627–1645. doi: 10.1517/14712598.7.11.1627. [DOI] [PubMed] [Google Scholar]
  • 102.Zheng X, Zhang X, Sun H, Feng B, Li M, Chen G, Vladau C, Chen D, Suzuki M, Min L, Liu W, Zhong R, Garcia B, Jevnikar A, Min WP. Protection of renal ischemia injury using combination gene silencing of complement 3 and caspase 3 genes. Transplantation. 2006;82:1781–1786. doi: 10.1097/01.tp.0000250769.86623.a3. [DOI] [PubMed] [Google Scholar]
  • 103.Zhang X, Zheng X, Sun H, Feng B, Chen G, Vladau C, Li M, Chen D, Suzuki M, Min L, Liu W, Garcia B, Zhong R, Min WP. Prevention of renal ischemic injury by silencing the expression of renal caspase 3 and caspase 8. Transplantation. 2006;82:1728–1732. doi: 10.1097/01.tp.0000250764.17636.ba. [DOI] [PubMed] [Google Scholar]
  • 104.Contreras JL, Vilatoba M, Eckstein C, Bilbao G, Anthony Thompson J, Eckhoff DE. Caspase-8 and caspase-3 small interfering RNA decreases ischemia/reperfusion injury to the liver in mice. Surgery. 2004;136:390–400. doi: 10.1016/j.surg.2004.05.015. [DOI] [PubMed] [Google Scholar]
  • 105.Cheng G, Zhu L, Mahato RI. Caspase-3 gene silencing for inhibiting apoptosis in insulinoma cells and human islets. Mol Pharm. 2008;5:1093–1102. doi: 10.1021/mp800093f. [DOI] [PubMed] [Google Scholar]
  • 106.Burkhardt BR, Lyle R, Qian K, Arnold AS, Cheng H, Atkinson MA, Zhang YC. Efficient delivery of siRNA into cytokine-stimulated insulinoma cells silences Fas expression and inhibits Fas-mediated apoptosis. FEBS Lett. 2006;580:553–560. doi: 10.1016/j.febslet.2005.12.068. [DOI] [PubMed] [Google Scholar]
  • 107.Wang J, Li W, Min J, Ou Q, Chen J. Fas siRNA reduces apoptotic cell death of allogeneic-transplanted hepatocytes in mouse spleen. Transplant Proc. 2003;35:1594–1595. doi: 10.1016/s0041-1345(03)00438-x. [DOI] [PubMed] [Google Scholar]
  • 108.Li X, Zhang JF, Lu MQ, Yang Y, Xu C, Li H, Wang GS, Cai CJ, Chen GH. Alleviation of ischemia-reperfusion injury in rat liver transplantation by induction of small interference RNA targeting Fas. Langenbecks Arch Surg. 2007;392:345–351. doi: 10.1007/s00423-006-0142-5. [DOI] [PubMed] [Google Scholar]
  • 109.Zheng X, Lian D, Wong A, Bygrave M, Ichim TE, Khoshniat M, Zhang X, Sun H, De Zordo T, Lacefield JC, Garcia B, Jevnikar AM, Min WP. Novel small interfering RNA-containing solution protecting donor organs in heart transplantation. Circulation. 2009;120:1099–1107. doi: 10.1161/CIRCULATIONAHA.108.787390. 1091 p following 1107. [DOI] [PubMed] [Google Scholar]
  • 110.Zheng X, Zhang X, Feng B, Sun H, Suzuki M, Ichim T, Kubo N, Wong A, Min LR, Budohn ME, Garcia B, Jevnikar AM, Min WP. Gene silencing of complement C5a receptor using siRNA for preventing ischemia/reperfusion injury. Am J Pathol. 2008;173:973–980. doi: 10.2353/ajpath.2008.080103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.McCabe C, O'Brien T. Beta cell cytoprotection using lentiviral vector-based iNOS-specific shRNA delivery. Biochem Biophys Res Commun. 2007;357:75–80. doi: 10.1016/j.bbrc.2007.03.115. [DOI] [PubMed] [Google Scholar]
  • 112.McCabe C, O'Brien T. The rational design of beta cell cytoprotective gene transfer strategies: targeting deleterious iNOS expression. Mol Biotechnol. 2007;37:38–47. doi: 10.1007/s12033-007-0049-6. [DOI] [PubMed] [Google Scholar]
  • 113.Li F, Mahato RI. iNOS gene silencing prevents inflammatory cytokine-induced beta-cell apoptosis. Mol Pharm. 2008;5:407–417. doi: 10.1021/mp700145f. [DOI] [PubMed] [Google Scholar]
  • 114.De Paula D, Bentley MV, Mahato RI. Effect of iNOS and NF-kappaB gene silencing on beta-cell survival and function. J Drug Target. 2007;15:358–369. doi: 10.1080/10611860701349695. [DOI] [PubMed] [Google Scholar]
  • 115.Li F, Mahato RI. Bipartite vectors for co-expression of a growth factor cDNA and short hairpin RNA against an apoptotic gene. J Gene Med. 2009;11:764–771. doi: 10.1002/jgm.1357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Pinkenburg O, Platz J, Beisswenger C, Vogelmeier C, Bals R. Inhibition of NF-kappaB mediated inflammation by siRNA expressed by recombinant adeno-associated virus. J Virol Methods. 2004;120:119–122. doi: 10.1016/j.jviromet.2004.04.007. [DOI] [PubMed] [Google Scholar]
  • 117.Feng B, Chen G, Zheng X, Sun H, Zhang X, Zhang ZX, Xiang Y, Ichim TE, Garcia B, Luke P, Jevnikar AM, Min WP. Small interfering RNA targeting RelB protects against renal ischemia-reperfusion injury. Transplantation. 2009;87:1283–1289. doi: 10.1097/TP.0b013e3181a1905e. [DOI] [PubMed] [Google Scholar]
  • 118.Lee YS, Nakahara K, Pham JW, Kim K, He Z, Sontheimer EJ, Carthew RW. Distinct roles for Drosophila Dicer-1 and Dicer-2 in the siRNA/miRNA silencing pathways. Cell. 2004;117:69–81. doi: 10.1016/s0092-8674(04)00261-2. [DOI] [PubMed] [Google Scholar]
  • 119.Wang L, Mu FY. A Web-based design center for vector-based siRNA and siRNA cassette. Bioinformatics. 2004;20:1818–1820. doi: 10.1093/bioinformatics/bth164. [DOI] [PubMed] [Google Scholar]
  • 120.Naito Y, Yamada T, Ui-Tei K, Morishita S, Saigo K. siDirect: highly effective, target-specific siRNA design software for mammalian RNA interference. Nucleic Acids Res. 2004;32:W124–W129. doi: 10.1093/nar/gkh442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Yuan B, Latek R, Hossbach M, Tuschl T, Lewitter F. siRNA Selection Server: an automated siRNA oligonucleotide prediction server. Nucleic Acids Res. 2004;32:W130–W134. doi: 10.1093/nar/gkh366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Henschel A, Buchholz F, Habermann B. DEQOR: a web-based tool for the design and quality control of siRNAs. Nucleic Acids Res. 2004;32:W113–W120. doi: 10.1093/nar/gkh408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Li L, Lin X, Khvorova A, Fesik SW, Shen Y. Defining the optimal parameters for hairpin-based knockdown constructs. Rna. 2007;13:1765–1774. doi: 10.1261/rna.599107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Harborth J, Elbashir SM, Vandenburgh K, Manninga H, Scaringe SA, Weber K, Tuschl T. Sequence, chemical, and structural variation of small interfering RNAs and short hairpin RNAs and the effect on mammalian gene silencing. Antisense Nucleic Acid Drug Dev. 2003;13:83–105. doi: 10.1089/108729003321629638. [DOI] [PubMed] [Google Scholar]
  • 125.Jackson AL, Burchard J, Schelter J, Chau BN, Cleary M, Lim L, Linsley PS. Widespread siRNA "off-target" transcript silencing mediated by seed region sequence complementarity. Rna. 2006;12:1179–1187. doi: 10.1261/rna.25706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Judge AD, Sood V, Shaw JR, Fang D, McClintock K, MacLachlan I. Sequence-dependent stimulation of the mammalian innate immune response by synthetic siRNA. Nat Biotechnol. 2005;23:457–462. doi: 10.1038/nbt1081. [DOI] [PubMed] [Google Scholar]
  • 127.Taxman DJ, Livingstone LR, Zhang J, Conti BJ, Iocca HA, Williams KL, Lich JD, Ting JP, Reed W. Criteria for effective design, construction, and gene knockdown by shRNA vectors. BMC Biotechnol. 2006;6:7. doi: 10.1186/1472-6750-6-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Hernandez N. Small nuclear RNA genes: a model system to study fundamental mechanisms of transcription. J Biol Chem. 2001;276:26733–26736. doi: 10.1074/jbc.R100032200. [DOI] [PubMed] [Google Scholar]
  • 129.Song J, Pang S, Lu Y, Chiu R. Poly(U) and polyadenylation termination signals are interchangeable for terminating the expression of shRNA from a pol II promoter. Biochem Biophys Res Commun. 2004;323:573–578. doi: 10.1016/j.bbrc.2004.08.128. [DOI] [PubMed] [Google Scholar]
  • 130.Zeng Y, Wagner EJ, Cullen BR. Both natural and designed micro RNAs can inhibit the expression of cognate mRNAs when expressed in human cells. Mol Cell. 2002;9:1327–1333. doi: 10.1016/s1097-2765(02)00541-5. [DOI] [PubMed] [Google Scholar]
  • 131.Denti MA, Rosa A, Sthandier O, De Angelis FG, Bozzoni I. A new vector, based on the PolII promoter of the U1 snRNA gene, for the expression of siRNAs in mammalian cells. Mol Ther. 2004;10:191–199. doi: 10.1016/j.ymthe.2004.04.008. [DOI] [PubMed] [Google Scholar]
  • 132.Huang M, Jia FJ, Yan YC, Guo LH, Li YP. Transactivated minimal E1b promoter is capable of driving the expression of short hairpin RNA. J Virol Methods. 2006;134:48–54. doi: 10.1016/j.jviromet.2005.11.016. [DOI] [PubMed] [Google Scholar]
  • 133.Xia XG, Zhou H, Ding H, Affar el B, Shi Y, Xu Z. An enhanced U6 promoter for synthesis of short hairpin RNA. Nucleic Acids Res. 2003;31:e100. doi: 10.1093/nar/gng098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Ong ST, Li F, Du J, Tan YW, Wang S. Hybrid cytomegalovirus enhancer-h1 promoter-based plasmid and baculovirus vectors mediate effective RNA interference. Hum Gene Ther. 2005;16:1404–1412. doi: 10.1089/hum.2005.16.1404. [DOI] [PubMed] [Google Scholar]
  • 135.Hassani Z, Francois JC, Alfama G, Dubois GM, Paris M, Giovannangeli C, Demeneix BA. A hybrid CMV-H1 construct improves efficiency of PEI-delivered shRNA in the mouse brain. Nucleic Acids Res. 2007;35:e65. doi: 10.1093/nar/gkm152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Brenz Verca MS, Weber P, Mayer C, Graf C, Refojo D, Kuhn R, Grummt I, Lutz B. Development of a species-specific RNA polymerase I-based shRNA expression vector. Nucleic Acids Res. 2007;35:e10. doi: 10.1093/nar/gkl1045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Lee SK, Kumar P. Conditional RNAi: towards a silent gene therapy. Adv Drug Deliv Rev. 2009;61:650–664. doi: 10.1016/j.addr.2009.03.016. [DOI] [PubMed] [Google Scholar]
  • 138.Czauderna F, Santel A, Hinz M, Fechtner M, Durieux B, Fisch G, Leenders F, Arnold W, Giese K, Klippel A, Kaufmann J. Inducible shRNA expression for application in a prostate cancer mouse model. Nucleic Acids Res. 2003;31:e127. doi: 10.1093/nar/gng127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.van de Wetering M, Oving I, Muncan V, Pon Fong MT, Brantjes H, van Leenen D, Holstege FC, Brummelkamp TR, Agami R, Clevers H. Specific inhibition of gene expression using a stably integrated, inducible small-interfering-RNA vector. EMBO Rep. 2003;4:609–615. doi: 10.1038/sj.embor.embor865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Higuchi M, Tsutsumi R, Higashi H, Hatakeyama M. Conditional gene silencing utilizing the lac repressor reveals a role of SHP-2 in cagA-positive Helicobacter pylori pathogenicity. Cancer Sci. 2004;95:442–447. doi: 10.1111/j.1349-7006.2004.tb03229.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Heinonen JE, Mohamed AJ, Nore BF, Smith CI. Inducible H1 promoter-driven lentiviral siRNA expression by Stuffer reporter deletion. Oligonucleotides. 2005;15:139–144. doi: 10.1089/oli.2005.15.139. [DOI] [PubMed] [Google Scholar]
  • 142.Kim HA, Lee BW, Kang D, Kim JH, Ihm SH, Lee M. Delivery of hypoxia-inducible VEGF gene to rat islets using polyethylenimine. J Drug Target. 2009;17:1–9. doi: 10.1080/10611860802392982. [DOI] [PubMed] [Google Scholar]
  • 143.Lee BW, Chae HY, Tuyen TT, Kang D, Kim HA, Lee M, Ihm SH. A comparison of non-viral vectors for gene delivery to pancreatic beta-cells: delivering a hypoxia-inducible vascular endothelial growth factor gene to rat islets. Int J Mol Med. 2009;23:757–762. doi: 10.3892/ijmm_00000189. [DOI] [PubMed] [Google Scholar]
  • 144.Chai R, Chen S, Ding J, Grayburn PA. Efficient, glucose responsive and islet-specific transgene expression by a modified rat insulin promoter. Gene Ther. 2009;16:1202–1209. doi: 10.1038/gt.2009.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Iwakura H, Hosoda K, Son C, Fujikura J, Tomita T, Noguchi M, Ariyasu H, Takaya K, Masuzaki H, Ogawa Y, Hayashi T, Inoue G, Akamizu T, Hosoda H, Kojima M, Itoh H, Toyokuni S, Kangawa K, Nakao K. Analysis of rat insulin II promoter-ghrelin transgenic mice and rat glucagon promoter-ghrelin transgenic mice. J Biol Chem. 2005;280:15247–15256. doi: 10.1074/jbc.M411358200. [DOI] [PubMed] [Google Scholar]
  • 146.Grzech M, Dahlhoff M, Herbach N, Habermann FA, Renner-Muller I, Wanke R, Flaswinkel H, Wolf E, Schneider MR. Specific transgene expression in mouse pancreatic beta-cells under the control of the porcine insulin promoter. Mol Cell Endocrinol. 315:219–224. doi: 10.1016/j.mce.2009.08.001. [DOI] [PubMed] [Google Scholar]
  • 147.Avnit-Sagi T, Kantorovich L, Kredo-Russo S, Hornstein E, Walker MD. The promoter of the pri-miR-375 gene directs expression selectively to the endocrine pancreas. PLoS One. 2009;4:e5033. doi: 10.1371/journal.pone.0005033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Siolas D, Lerner C, Burchard J, Ge W, Linsley PS, Paddison PJ, Hannon GJ, Cleary MA. Synthetic shRNAs as potent RNAi triggers. Nat Biotechnol. 2005;23:227–231. doi: 10.1038/nbt1052. [DOI] [PubMed] [Google Scholar]
  • 149.Jeanson-Leh L, Blondeau J, Galy A. Optimization of short hairpin RNA for lentiviral-mediated RNAi against WAS. Biochem Biophys Res Commun. 2007;362:498–503. doi: 10.1016/j.bbrc.2007.08.013. [DOI] [PubMed] [Google Scholar]
  • 150.Boden D, Pusch O, Silbermann R, Lee F, Tucker L, Ramratnam B. Enhanced gene silencing of HIV-1 specific siRNA using microRNA designed hairpins. Nucleic Acids Res. 2004;32:1154–1158. doi: 10.1093/nar/gkh278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Silva JM, Li MZ, Chang K, Ge W, Golding MC, Rickles RJ, Siolas D, Hu G, Paddison PJ, Schlabach MR, Sheth N, Bradshaw J, Burchard J, Kulkarni A, Cavet G, Sachidanandam R, McCombie WR, Cleary MA, Elledge SJ, Hannon GJ. Second-generation shRNA libraries covering the mouse and human genomes. Nat Genet. 2005;37:1281–1288. doi: 10.1038/ng1650. [DOI] [PubMed] [Google Scholar]
  • 152.Stegmeier F, Hu G, Rickles RJ, Hannon GJ, Elledge SJ. A lentiviral microRNA-based system for single-copy polymerase II-regulated RNA interference in mammalian cells. Proc Natl Acad Sci U S A. 2005;102:13212–13217. doi: 10.1073/pnas.0506306102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Boudreau RL, Monteys AM, Davidson BL. Minimizing variables among hairpin-based RNAi vectors reveals the potency of shRNAs. Rna. 2008;14:1834–1844. doi: 10.1261/rna.1062908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Boudreau RL, Martins I, Davidson BL. Artificial microRNAs as siRNA shuttles: improved safety as compared to shRNAs in vitro and in vivo. Mol Ther. 2009;17:169–175. doi: 10.1038/mt.2008.231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Nielsen TT, Marion I, Hasholt L, Lundberg C. Neuron-specific RNA interference using lentiviral vectors. J Gene Med. 2009;11:559–569. doi: 10.1002/jgm.1333. [DOI] [PubMed] [Google Scholar]
  • 156.Chung KH, Hart CC, Al-Bassam S, Avery A, Taylor J, Patel PD, Vojtek AB, Turner DL. Polycistronic RNA polymerase II expression vectors for RNA interference based on BIC/miR-155. Nucleic Acids Res. 2006;34:e53. doi: 10.1093/nar/gkl143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Aagaard LA, Zhang J, von Eije KJ, Li H, Saetrom P, Amarzguioui M, Rossi JJ. Engineering and optimization of the miR-106b cluster for ectopic expression of multiplexed anti-HIV RNAs. Gene Ther. 2008;15:1536–1549. doi: 10.1038/gt.2008.147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Scherer LJ, Frank R, Rossi JJ. Optimization and characterization of tRNA-shRNA expression constructs. Nucleic Acids Res. 2007;35:2620–2628. doi: 10.1093/nar/gkm103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Kawasaki H, Taira K. Short hairpin type of dsRNAs that are controlled by tRNA(Val) promoter significantly induce RNAi-mediated gene silencing in the cytoplasm of human cells. Nucleic Acids Res. 2003;31:700–707. doi: 10.1093/nar/gkg158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Boden D, Pusch O, Lee F, Tucker L, Shank PR, Ramratnam B. Promoter choice affects the potency of HIV-1 specific RNA interference. Nucleic Acids Res. 2003;31:5033–5038. doi: 10.1093/nar/gkg704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Kuwabara T, Warashina M, Sano M, Tang H, Wong-Staal F, Munekata E, Taira K. Recognition of engineered tRNAs with an extended 3' end by Exportin-t (Xpo-t) and transport of tRNA-attached ribozymes to the cytoplasm in somatic cells. Biomacromolecules. 2001;2:1229–1242. doi: 10.1021/bm0101062. [DOI] [PubMed] [Google Scholar]
  • 162.Zheng L, Liu J, Batalov S, Zhou D, Orth A, Ding S, Schultz PG. An approach to genomewide screens of expressed small interfering RNAs in mammalian cells. Proc Natl Acad Sci U S A. 2004;101:135–140. doi: 10.1073/pnas.2136685100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Jackson AL, Bartz SR, Schelter J, Kobayashi SV, Burchard J, Mao M, Li B, Cavet G, Linsley PS. Expression profiling reveals off-target gene regulation by RNAi. Nat Biotechnol. 2003;21:635–637. doi: 10.1038/nbt831. [DOI] [PubMed] [Google Scholar]
  • 164.Jackson AL, Linsley PS. Recognizing and avoiding siRNA off-target effects for target identification and therapeutic application. Nat Rev Drug Discov. 2010;9:57–67. doi: 10.1038/nrd3010. [DOI] [PubMed] [Google Scholar]
  • 165.Rao DD, Vorhies JS, Senzer N, Nemunaitis J. siRNA vs. shRNA: similarities and differences. Adv Drug Deliv Rev. 2009;61:746–759. doi: 10.1016/j.addr.2009.04.004. [DOI] [PubMed] [Google Scholar]
  • 166.Jackson AL, Burchard J, Leake D, Reynolds A, Schelter J, Guo J, Johnson JM, Lim L, Karpilow J, Nichols K, Marshall W, Khvorova A, Linsley PS. Position-specific chemical modification of siRNAs reduces "off-target" transcript silencing. Rna. 2006;12:1197–1205. doi: 10.1261/rna.30706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Lin X, Ruan X, Anderson MG, McDowell JA, Kroeger PE, Fesik SW, Shen Y. siRNA-mediated off-target gene silencing triggered by a 7 nt complementation. Nucleic Acids Res. 2005;33:4527–4535. doi: 10.1093/nar/gki762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Qiu S, Adema CM, Lane T. A computational study of off-target effects of RNA interference. Nucleic Acids Res. 2005;33:1834–1847. doi: 10.1093/nar/gki324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Du Q, Thonberg H, Wang J, Wahlestedt C, Liang Z. A systematic analysis of the silencing effects of an active siRNA at all single-nucleotide mismatched target sites. Nucleic Acids Res. 2005;33:1671–1677. doi: 10.1093/nar/gki312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Bramsen JB, Laursen MB, Damgaard CK, Lena SW, Babu BR, Wengel J, Kjems J. Improved silencing properties using small internally segmented interfering RNAs. Nucleic Acids Res. 2007;35:5886–5897. doi: 10.1093/nar/gkm548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Ui-Tei K, Naito Y, Zenno S, Nishi K, Yamato K, Takahashi F, Juni A, Saigo K. Functional dissection of siRNA sequence by systematic DNA substitution: modified siRNA with a DNA seed arm is a powerful tool for mammalian gene silencing with significantly reduced off-target effect. Nucleic Acids Res. 2008;36:2136–2151. doi: 10.1093/nar/gkn042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Marques JT, Devosse T, Wang D, Zamanian-Daryoush M, Serbinowski P, Hartmann R, Fujita T, Behlke MA, Williams BR. A structural basis for discriminating between self and nonself double-stranded RNAs in mammalian cells. Nat Biotechnol. 2006;24:559–565. doi: 10.1038/nbt1205. [DOI] [PubMed] [Google Scholar]
  • 173.Kim DH, Behlke MA, Rose SD, Chang MS, Choi S, Rossi JJ. Synthetic dsRNA Dicer substrates enhance RNAi potency and efficacy. Nat Biotechnol. 2005;23:222–226. doi: 10.1038/nbt1051. [DOI] [PubMed] [Google Scholar]
  • 174.Marques JT, Williams BR. Activation of the mammalian immune system by siRNAs. Nat Biotechnol. 2005;23:1399–1405. doi: 10.1038/nbt1161. [DOI] [PubMed] [Google Scholar]
  • 175.Saunders LR, Barber GN. The dsRNA binding protein family: critical roles, diverse cellular functions. FASEB J. 2003;17:961–983. doi: 10.1096/fj.02-0958rev. [DOI] [PubMed] [Google Scholar]
  • 176.Hornung V, Guenthner-Biller M, Bourquin C, Ablasser A, Schlee M, Uematsu S, Noronha A, Manoharan M, Akira S, de Fougerolles A, Endres S, Hartmann G. Sequence-specific potent induction of IFN-alpha by short interfering RNA in plasmacytoid dendritic cells through TLR7. Nat Med. 2005;11:263–270. doi: 10.1038/nm1191. [DOI] [PubMed] [Google Scholar]
  • 177.Sledz CA, Holko M, de Veer MJ, Silverman RH, Williams BR. Activation of the interferon system by short-interfering RNAs. Nat Cell Biol. 2003;5:834–839. doi: 10.1038/ncb1038. [DOI] [PubMed] [Google Scholar]
  • 178.Hemmi H, Takeuchi O, Kawai T, Kaisho T, Sato S, Sanjo H, Matsumoto M, Hoshino K, Wagner H, Takeda K, Akira S. A Toll-like receptor recognizes bacterial DNA. Nature. 2000;408:740–745. doi: 10.1038/35047123. [DOI] [PubMed] [Google Scholar]
  • 179.de Wolf HK, Johansson N, Thong AT, Snel CJ, Mastrobattista E, Hennink WE, Storm G. Plasmid CpG depletion improves degree and duration of tumor gene expression after intravenous administration of polyplexes. Pharm Res. 2008;25:1654–1662. doi: 10.1007/s11095-008-9558-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Grimm D, Kay MA. Therapeutic application of RNAi: is mRNA targeting finally ready for prime time? J Clin Invest. 2007;117:3633–3641. doi: 10.1172/JCI34129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Cheng K, Yang N, Mahato RI. TGF-beta1 gene silencing for treating liver fibrosis. Mol Pharm. 2009;6:772–779. doi: 10.1021/mp9000469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Chen Y, Mahato RI. siRNA pool targeting different sites of human hepatitis B surface antigen efficiently inhibits HBV infection. J Drug Target. 2008;16:140–148. doi: 10.1080/10611860701878750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Grimm D, Kay MA. Combinatorial RNAi: a winning strategy for the race against evolving targets? Mol Ther. 2007;15:878–888. doi: 10.1038/sj.mt.6300116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Gonzalez S, Castanotto D, Li H, Olivares S, Jensen MC, Forman SJ, Rossi JJ, Cooper LJ. Amplification of RNAi--targeting HLA mRNAs. Mol Ther. 2005;11:811–818. doi: 10.1016/j.ymthe.2004.12.023. [DOI] [PubMed] [Google Scholar]
  • 185.Zhou H, Xia XG, Xu Z. An RNA polymerase II construct synthesizes short-hairpin RNA with a quantitative indicator and mediates highly efficient RNAi. Nucleic Acids Res. 2005;33:e62. doi: 10.1093/nar/gni061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Xia XG, Zhou H, Samper E, Melov S, Xu Z. Pol II-expressed shRNA knocks down Sod2 gene expression and causes phenotypes of the gene knockout in mice. PLoS Genet. 2006;2:e10. doi: 10.1371/journal.pgen.0020010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Sun D, Melegari M, Sridhar S, Rogler CE, Zhu L. Multi-miRNA hairpin method that improves gene knockdown efficiency and provides linked multi-gene knockdown. Biotechniques. 2006;41:59–63. doi: 10.2144/000112203. [DOI] [PubMed] [Google Scholar]
  • 188.Weinberg MS, Ely A, Barichievy S, Crowther C, Mufamadi S, Carmona S, Arbuthnot P. Specific inhibition of HBV replication in vitro and in vivo with expressed long hairpin RNA. Mol Ther. 2007;15:534–541. doi: 10.1038/sj.mt.6300077. [DOI] [PubMed] [Google Scholar]
  • 189.Saayman S, Barichievy S, Capovilla A, Morris KV, Arbuthnot P, Weinberg MS. The efficacy of generating three independent anti-HIV-1 siRNAs from a single U6 RNA Pol III-expressed long hairpin RNA. PLoS One. 2008;3:e2602. doi: 10.1371/journal.pone.0002602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Liu YP, Haasnoot J, Berkhout B. Design of extended short hairpin RNAs for HIV-1 inhibition. Nucleic Acids Res. 2007;35:5683–5693. doi: 10.1093/nar/gkm596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Liu YP, von Eije KJ, Schopman NC, Westerink JT, ter Brake O, Haasnoot J, Berkhout B. Combinatorial RNAi against HIV-1 using extended short hairpin RNAs. Mol Ther. 2009;17:1712–1723. doi: 10.1038/mt.2009.176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Yu JY, Taylor J, DeRuiter SL, Vojtek AB, Turner DL. Simultaneous inhibition of GSK3alpha and GSK3beta using hairpin siRNA expression vectors. Mol Ther. 2003;7:228–236. doi: 10.1016/s1525-0016(02)00037-0. [DOI] [PubMed] [Google Scholar]
  • 193.Jazag A, Kanai F, Ijichi H, Tateishi K, Ikenoue T, Tanaka Y, Ohta M, Imamura J, Guleng B, Asaoka Y, Kawabe T, Miyagishi M, Taira K, Omata M. Single small-interfering RNA expression vector for silencing multiple transforming growth factor-beta pathway components. Nucleic Acids Res. 2005;33:e131. doi: 10.1093/nar/gni130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Shin KJ, Wall EA, Zavzavadjian JR, Santat LA, Liu J, Hwang JI, Rebres R, Roach T, Seaman W, Simon MI, Fraser ID. A single lentiviral vector platform for microRNA-based conditional RNA interference and coordinated transgene expression. Proc Natl Acad Sci U S A. 2006;103:13759–13764. doi: 10.1073/pnas.0606179103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Samakoglu S, Lisowski L, Budak-Alpdogan T, Usachenko Y, Acuto S, Di Marzo R, Maggio A, Zhu P, Tisdale JF, Riviere I, Sadelain M. A genetic strategy to treat sickle cell anemia by coregulating globin transgene expression and RNA interference. Nat Biotechnol. 2006;24:89–94. doi: 10.1038/nbt1176. [DOI] [PubMed] [Google Scholar]
  • 196.Qiu L, Wang H, Xia X, Zhou H, Xu Z. A construct with fluorescent indicators for conditional expression of miRNA. BMC Biotechnol. 2008;8:77. doi: 10.1186/1472-6750-8-77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Juliano R, Bauman J, Kang H, Ming X. Biological Barriers to Therapy with Antisense and siRNA Oligonucleotides. Mol Pharm. 2009;6:686–695. doi: 10.1021/mp900093r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Lewis DL, Wolff JA. Systemic siRNA delivery via hydrodynamic intravascular injection. Adv Drug Deliv Rev. 2007;59:115–123. doi: 10.1016/j.addr.2007.03.002. [DOI] [PubMed] [Google Scholar]
  • 199.Bradley SP, Kowalik TF, Rastellini C, da Costa MA, Bloomenthal AB, Cicalese L, Basadonna GP, Uknis ME. Successful incorporation of short-interfering RNA into islet cells by in situ perfusion. Transplant Proc. 2005;37:233–236. doi: 10.1016/j.transproceed.2004.12.181. [DOI] [PubMed] [Google Scholar]
  • 200.Lefebvre B, Vandewalle B, Longue J, Moerman E, Lukowiak B, Gmyr V, Maedler K, Kerr-conte J, Pattou F. Efficient gene delivery and silencing of mouse and human pancreatic islets. BMC Biotechnol. 10:28. doi: 10.1186/1472-6750-10-28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Behlke MA. Chemical modification of siRNAs for in vivo use. Oligonucleotides. 2008;18:305–319. doi: 10.1089/oli.2008.0164. [DOI] [PubMed] [Google Scholar]
  • 202.Corey DR. Chemical modification: the key to clinical application of RNA interference? J Clin Invest. 2007;117:3615–3622. doi: 10.1172/JCI33483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Morrissey DV, Blanchard K, Shaw L, Jensen K, Lockridge JA, Dickinson B, McSwiggen JA, Vargeese C, Bowman K, Shaffer CS, Polisky BA, Zinnen S. Activity of stabilized short interfering RNA in a mouse model of hepatitis B virus replication. Hepatology. 2005;41:1349–1356. doi: 10.1002/hep.20702. [DOI] [PubMed] [Google Scholar]
  • 204.Morrissey DV, Lockridge JA, Shaw L, Blanchard K, Jensen K, Breen W, Hartsough K, Machemer L, Radka S, Jadhav V, Vaish N, Zinnen S, Vargeese C, Bowman K, Shaffer CS, Jeffs LB, Judge A, MacLachlan I, Polisky B. Potent and persistent in vivo anti-HBV activity of chemically modified siRNAs. Nat Biotechnol. 2005;23:1002–1007. doi: 10.1038/nbt1122. [DOI] [PubMed] [Google Scholar]
  • 205.Choung S, Kim YJ, Kim S, Park HO, Choi YC. Chemical modification of siRNAs to improve serum stability without loss of efficacy. Biochem Biophys Res Commun. 2006;342:919–927. doi: 10.1016/j.bbrc.2006.02.049. [DOI] [PubMed] [Google Scholar]
  • 206.Czauderna F, Fechtner M, Dames S, Aygun H, Klippel A, Pronk GJ, Giese K, Kaufmann J. Structural variations and stabilising modifications of synthetic siRNAs in mammalian cells. Nucleic Acids Res. 2003;31:2705–2716. doi: 10.1093/nar/gkg393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Santel A, Aleku M, Keil O, Endruschat J, Esche V, Durieux B, Loffler K, Fechtner M, Rohl T, Fisch G, Dames S, Arnold W, Giese K, Klippel A, Kaufmann J. RNA interference in the mouse vascular endothelium by systemic administration of siRNA-lipoplexes for cancer therapy. Gene Ther. 2006;13:1360–1370. doi: 10.1038/sj.gt.3302778. [DOI] [PubMed] [Google Scholar]
  • 208.Turner JJ, Jones SW, Moschos SA, Lindsay MA, Gait MJ. MALDI-TOF mass spectral analysis of siRNA degradation in serum confirms an RNAse A-like activity. Mol Biosyst. 2007;3:43–50. doi: 10.1039/b611612d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Judge AD, Bola G, Lee AC, MacLachlan I. Design of noninflammatory synthetic siRNA mediating potent gene silencing in vivo. Mol Ther. 2006;13:494–505. doi: 10.1016/j.ymthe.2005.11.002. [DOI] [PubMed] [Google Scholar]
  • 210.Robbins M, Judge A, Liang L, McClintock K, Yaworski E, MacLachlan I. 2'-Omethyl-modified RNAs act as TLR7 antagonists. Mol Ther. 2007;15:1663–1669. doi: 10.1038/sj.mt.6300240. [DOI] [PubMed] [Google Scholar]
  • 211.Felgner PL, Gadek TR, Holm M, Roman R, Chan HW, Wenz M, Northrop JP, Ringold GM, Danielsen M. Lipofection: a highly efficient, lipid-mediated DNA-transfection procedure. Proc Natl Acad Sci U S A. 1987;84:7413–7417. doi: 10.1073/pnas.84.21.7413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Narang AS, Thoma L, Miller DD, Mahato RI. Cationic lipids with increased DNA binding affinity for nonviral gene transfer in dividing and nondividing cells. Bioconjug Chem. 2005;16:156–168. doi: 10.1021/bc049818q. [DOI] [PubMed] [Google Scholar]
  • 213.Takahashi Y, Nishikawa M, Takakura Y. Nonviral vector-mediated RNA interference: its gene silencing characteristics and important factors to achieve RNAi-based gene therapy. Adv Drug Deliv Rev. 2009;61:760–766. doi: 10.1016/j.addr.2009.04.006. [DOI] [PubMed] [Google Scholar]
  • 214.Breunig M, Lungwitz U, Liebl R, Goepferich A. Breaking up the correlation between efficacy and toxicity for nonviral gene delivery. Proc Natl Acad Sci U S A. 2007;104:14454–14459. doi: 10.1073/pnas.0703882104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Lee Y, Mo H, Koo H, Park JY, Cho MY, Jin GW, Park JS. Visualization of the degradation of a disulfide polymer, linear poly(ethylenimine sulfide), for gene delivery. Bioconjug Chem. 2007;18:13–18. doi: 10.1021/bc060113t. [DOI] [PubMed] [Google Scholar]
  • 216.Hoon Jeong J, Christensen LV, Yockman JW, Zhong Z, Engbersen JF, Jong Kim W, Feijen J, Wan Kim S. Reducible poly(amido ethylenimine) directed to enhance RNA interference. Biomaterials. 2007;28:1912–1917. doi: 10.1016/j.biomaterials.2006.12.019. [DOI] [PubMed] [Google Scholar]
  • 217.Breunig M, Hozsa C, Lungwitz U, Watanabe K, Umeda I, Kato H, Goepferich A. Mechanistic investigation of poly(ethylene imine)-based siRNA delivery: disulfide bonds boost intracellular release of the cargo. J Control Release. 2008;130:57–63. doi: 10.1016/j.jconrel.2008.05.016. [DOI] [PubMed] [Google Scholar]
  • 218.Stevenson M, Ramos-Perez V, Singh S, Soliman M, Preece JA, Briggs SS, Read ML, Seymour LW. Delivery of siRNA mediated by histidine-containing reducible polycations. J Control Release. 2008;130:46–56. doi: 10.1016/j.jconrel.2008.05.014. [DOI] [PubMed] [Google Scholar]
  • 219.Wang XL, Jensen R, Lu ZR. A novel environment-sensitive biodegradable polydisulfide with protonatable pendants for nucleic acid delivery. J Control Release. 2007;120:250–258. doi: 10.1016/j.jconrel.2007.05.014. [DOI] [PubMed] [Google Scholar]
  • 220.Ou M, Wang XL, Xu R, Chang CW, Bull DA, Kim SW. Novel biodegradable poly(disulfide amine)s for gene delivery with high efficiency and low cytotoxicity. Bioconjug Chem. 2008;19:626–633. doi: 10.1021/bc700397x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Green JJ, Shi J, Chiu E, Leshchiner ES, Langer R, Anderson DG. Biodegradable polymeric vectors for gene delivery to human endothelial cells. Bioconjug Chem. 2006;17:1162–1169. doi: 10.1021/bc0600968. [DOI] [PubMed] [Google Scholar]
  • 222.Jere D, Xu CX, Arote R, Yun CH, Cho MH, Cho CS. Poly(beta-amino ester) as a carrier for si/shRNA delivery in lung cancer cells. Biomaterials. 2008;29:2535–2547. doi: 10.1016/j.biomaterials.2008.02.018. [DOI] [PubMed] [Google Scholar]
  • 223.Vandenbroucke RE, De Geest BG, Bonne S, Vinken M, Van Haecke T, Heimberg H, Wagner E, Rogiers V, De Smedt SC, Demeester J, Sanders NN. Prolonged gene silencing in hepatoma cells and primary hepatocytes after small interfering RNA delivery with biodegradable poly(beta-amino esters) J Gene Med. 2008;10:783–794. doi: 10.1002/jgm.1202. [DOI] [PubMed] [Google Scholar]
  • 224.Tseng SJ, Tang SC. Development of poly(amino ester glycol urethane)/siRNA polyplexes for gene silencing. Bioconjug Chem. 2007;18:1383–1390. doi: 10.1021/bc060382j. [DOI] [PubMed] [Google Scholar]
  • 225.Zhao Z, Wang J, Mao HQ, Leong KW. Polyphosphoesters in drug and gene delivery. Adv Drug Deliv Rev. 2003;55:483–499. doi: 10.1016/s0169-409x(03)00040-1. [DOI] [PubMed] [Google Scholar]
  • 226.Mao HQ, Leong KW. Design of polyphosphoester-DNA nanoparticles for non-viral gene delivery. Adv Genet. 2005;53:275–306. [PubMed] [Google Scholar]
  • 227.Wang J, Zhang PC, Mao HQ, Leong KW. Enhanced gene expression in mouse muscle by sustained release of plasmid DNA using PPE-EA as a carrier. Gene Ther. 2002;9:1254–1261. doi: 10.1038/sj.gt.3301794. [DOI] [PubMed] [Google Scholar]
  • 228.Sun TM, Du JZ, Yan LF, Mao HQ, Wang J. Self-assembled biodegradable micellar nanoparticles of amphiphilic and cationic block copolymer for siRNA delivery. Biomaterials. 2008 doi: 10.1016/j.biomaterials.2008.07.036. [DOI] [PubMed] [Google Scholar]
  • 229.DeRouchey J, Schmidt C, Walker GF, Koch C, Plank C, Wagner E, Radler JO. Monomolecular assembly of siRNA and poly(ethylene glycol)-peptide copolymers. Biomacromolecules. 2008;9:724–732. doi: 10.1021/bm7011482. [DOI] [PubMed] [Google Scholar]
  • 230.Shim MS, Kwon YJ. Controlled delivery of plasmid DNA and siRNA to intracellular targets using ketalized polyethylenimine. Biomacromolecules. 2008;9:444–455. doi: 10.1021/bm7007313. [DOI] [PubMed] [Google Scholar]
  • 231.Soutschek J, Akinc A, Bramlage B, Charisse K, Constien R, Donoghue M, Elbashir S, Geick A, Hadwiger P, Harborth J, John M, Kesavan V, Lavine G, Pandey RK, Racie T, Rajeev KG, Rohl I, Toudjarska I, Wang G, Wuschko S, Bumcrot D, Koteliansky V, Limmer S, Manoharan M, Vornlocher HP. Therapeutic silencing of an endogenous gene by systemic administration of modified siRNAs. Nature. 2004;432:173–178. doi: 10.1038/nature03121. [DOI] [PubMed] [Google Scholar]
  • 232.Wolfrum C, Shi S, Jayaprakash KN, Jayaraman M, Wang G, Pandey RK, Rajeev KG, Nakayama T, Charrise K, Ndungo EM, Zimmermann T, Koteliansky V, Manoharan M, Stoffel M. Mechanisms and optimization of in vivo delivery of lipophilic siRNAs. Nat Biotechnol. 2007;25:1149–1157. doi: 10.1038/nbt1339. [DOI] [PubMed] [Google Scholar]
  • 233.Lorenz C, Hadwiger P, John M, Vornlocher HP, Unverzagt C. Steroid and lipid conjugates of siRNAs to enhance cellular uptake and gene silencing in liver cells. Bioorg Med Chem Lett. 2004;14:4975–4977. doi: 10.1016/j.bmcl.2004.07.018. [DOI] [PubMed] [Google Scholar]
  • 234.Nishina K, Unno T, Uno Y, Kubodera T, Kanouchi T, Mizusawa H, Yokota T. Efficient in vivo delivery of siRNA to the liver by conjugation of alpha-tocopherol. Mol Ther. 2008;16:734–740. doi: 10.1038/mt.2008.14. [DOI] [PubMed] [Google Scholar]
  • 235.Chiu YL, Ali A, Chu CY, Cao H, Rana TM. Visualizing a correlation between siRNA localization, cellular uptake, and RNAi in living cells. Chem Biol. 2004;11:1165–1175. doi: 10.1016/j.chembiol.2004.06.006. [DOI] [PubMed] [Google Scholar]
  • 236.Moschos SA, Jones SW, Perry MM, Williams AE, Erjefalt JS, Turner JJ, Barnes PJ, Sproat BS, Gait MJ, Lindsay MA. Lung delivery studies using siRNA conjugated to TAT(48–60) and penetratin reveal peptide induced reduction in gene expression and induction of innate immunity. Bioconjug Chem. 2007;18:1450–1459. doi: 10.1021/bc070077d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Muratovska A, Eccles MR. Conjugate for efficient delivery of short interfering RNA (siRNA) into mammalian cells. FEBS Lett. 2004;558:63–68. doi: 10.1016/S0014-5793(03)01505-9. [DOI] [PubMed] [Google Scholar]
  • 238.Lee SH, Kim SH, Park TG. Intracellular siRNA delivery system using polyelectrolyte complex micelles prepared from VEGF siRNA-PEG conjugate and cationic fusogenic peptide. Biochem Biophys Res Commun. 2007;357:511–516. doi: 10.1016/j.bbrc.2007.03.185. [DOI] [PubMed] [Google Scholar]
  • 239.Kim SH, Jeong JH, Lee SH, Kim SW, Park TG. PEG conjugated VEGF siRNA for anti-angiogenic gene therapy. J Control Release. 2006;116:123–129. doi: 10.1016/j.jconrel.2006.05.023. [DOI] [PubMed] [Google Scholar]
  • 240.Kim SH, Jeong JH, Lee SH, Kim SW, Park TG. Local and systemic delivery of VEGF siRNA using polyelectrolyte complex micelles for effective treatment of cancer. J Control Release. 2008;129:107–116. doi: 10.1016/j.jconrel.2008.03.008. [DOI] [PubMed] [Google Scholar]
  • 241.Oishi M, Nagasaki Y, Itaka K, Nishiyama N, Kataoka K. Lactosylated poly(ethylene glycol)-siRNA conjugate through acid-labile beta-thiopropionate linkage to construct pH-sensitive polyion complex micelles achieving enhanced gene silencing in hepatoma cells. J Am Chem Soc. 2005;127:1624–1625. doi: 10.1021/ja044941d. [DOI] [PubMed] [Google Scholar]
  • 242.Oishi M, Nagasaki Y, Nishiyama N, Itaka K, Takagi M, Shimamoto A, Furuichi Y, Kataoka K. Enhanced growth inhibition of hepatic multicellular tumor spheroids by lactosylated poly(ethylene glycol)-siRNA conjugate formulated in PEGylated polyplexes. ChemMedChem. 2007;2:1290–1297. doi: 10.1002/cmdc.200700076. [DOI] [PubMed] [Google Scholar]
  • 243.Medarova Z, Kumar M, Ng SW, Yang J, Barteneva N, Evgenov NV, Petkova V, Moore A. Multifunctional magnetic nanocarriers for image-tagged SiRNA delivery to intact pancreatic islets. Transplantation. 2008;86:1170–1177. doi: 10.1097/TP.0b013e31818a81b2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Narang AS, Sabek O, Gaber AO, Mahato RI. Co-expression of vascular endothelial growth factor and interleukin-1 receptor antagonist improves human islet survival and function. Pharm Res. 2006;23:1970–1982. doi: 10.1007/s11095-006-9065-7. [DOI] [PubMed] [Google Scholar]
  • 245.Bain JR, Schisler JC, Takeuchi K, Newgard CB, Becker TC. An adenovirus vector for efficient RNA interference-mediated suppression of target genes in insulinoma cells and pancreatic islets of langerhans. Diabetes. 2004;53:2190–2194. doi: 10.2337/diabetes.53.9.2190. [DOI] [PubMed] [Google Scholar]
  • 246.Iezzi M, Eliasson L, Fukuda M, Wollheim CB. Adenovirus-mediated silencing of synaptotagmin 9 inhibits Ca2+dependent insulin secretion in islets. FEBS Lett. 2005;579:5241–5246. doi: 10.1016/j.febslet.2005.08.047. [DOI] [PubMed] [Google Scholar]
  • 247.Jensen MV, Joseph JW, Ilkayeva O, Burgess S, Lu D, Ronnebaum SM, Odegaard M, Becker TC, Sherry AD, Newgard CB. Compensatory responses to pyruvate carboxylase suppression in islet beta-cells. Preservation of glucose-stimulated insulin secretion. J Biol Chem. 2006;281:22342–22351. doi: 10.1074/jbc.M604350200. [DOI] [PubMed] [Google Scholar]
  • 248.Li ZL, Xu WJ, Tian PX, Ding XM, Tian XH, Feng XS, Hou J. [Recombinant adenovirus-mediated RNA silencing of tissue factor expression in human islet: an in vitro study] Nan Fang Yi Ke Da Xue Xue Bao. 2007;27:1299–1302. [PubMed] [Google Scholar]
  • 249.Marzban L, Tomas A, Becker TC, Rosenberg L, Oberholzer J, Fraser PE, Halban PA, Verchere CB. Small interfering RNA-mediated suppression of proislet amyloid polypeptide expression inhibits islet amyloid formation and enhances survival of human islets in culture. Diabetes. 2008;57:3045–3055. doi: 10.2337/db08-0485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Naldini L. Lentiviruses as gene transfer agents for delivery to non-dividing cells. Curr Opin Biotechnol. 1998;9:457–463. doi: 10.1016/s0958-1669(98)80029-3. [DOI] [PubMed] [Google Scholar]
  • 251.Burns JC, Friedmann T, Driever W, Burrascano M, Yee JK. 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 U S A. 1993;90:8033–8037. doi: 10.1073/pnas.90.17.8033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Kobinger GP, Deng S, Louboutin JP, Vatamaniuk M, Matschinsky F, Markmann JF, Raper SE, Wilson JM. Transduction of human islets with pseudotyped lentiviral vectors. Hum Gene Ther. 2004;15:211–219. doi: 10.1089/104303404772680010. [DOI] [PubMed] [Google Scholar]
  • 253.Fenjves ES, Ochoa MS, Cechin S, Gay-Rabinstein C, Perez-Alvarez I, Ichii H, Mendez A, Ricordi C, Curran MA. Protection of human pancreatic islets using a lentiviral vector expressing two genes: cFLIP and GFP. Cell Transplant. 2008;17:793–802. doi: 10.3727/096368908786516828. [DOI] [PubMed] [Google Scholar]
  • 254.Lu Y, Dang H, Middleton B, Zhang Z, Washburn L, Campbell-Thompson M, Atkinson MA, Gambhir SS, Tian J, Kaufman DL. Bioluminescent monitoring of islet graft survival after transplantation. Mol Ther. 2004;9:428–435. doi: 10.1016/j.ymthe.2004.01.008. [DOI] [PubMed] [Google Scholar]
  • 255.Park IK, Lasiene J, Chou SH, Horner PJ, Pun SH. Neuron-specific delivery of nucleic acids mediated by Tet1-modified poly(ethylenimine) J Gene Med. 2007;9:691–702. doi: 10.1002/jgm.1062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Fernandes JR, Duvivier-Kali VF, Keegan M, Hollister-Lock J, Omer A, Su S, Bonner-Weir S, Feng S, Lee JS, Mulligan RC, Weir GC. Transplantation of islets transduced with CTLA4-Ig and TGFbeta using adenovirus and lentivirus vectors. Transpl Immunol. 2004;13:191–200. doi: 10.1016/j.trim.2004.04.004. [DOI] [PubMed] [Google Scholar]
  • 257.Giannoukakis N, Mi Z, Gambotto A, Eramo A, Ricordi C, Trucco M, Robbins P. Infection of intact human islets by a lentiviral vector. Gene Ther. 1999;6:1545–1551. doi: 10.1038/sj.gt.3300996. [DOI] [PubMed] [Google Scholar]
  • 258.Gallichan WS, Kafri T, Krahl T, Verma IM, Sarvetnick N. Lentivirus-mediated transduction of islet grafts with interleukin 4 results in sustained gene expression and protection from insulitis. Hum Gene Ther. 1998;9:2717–2726. doi: 10.1089/hum.1998.9.18-2717. [DOI] [PubMed] [Google Scholar]
  • 259.He Z, Wang F, Kumagai-Braesch M, Permert J, Holgersson J. Long-term gene expression and metabolic control exerted by lentivirus-transduced pancreatic islets. Xenotransplantation. 2006;13:195–203. doi: 10.1111/j.1399-3089.2006.00274.x. [DOI] [PubMed] [Google Scholar]
  • 260.Callewaert H, Gysemans C, Cardozo AK, Elsner M, Tiedge M, Eizirik DL, Mathieu C. Cell loss during pseudoislet formation hampers profound improvements in islet lentiviral transduction efficacy for transplantation purposes. Cell Transplant. 2007;16:527–537. doi: 10.3727/000000007783464948. [DOI] [PubMed] [Google Scholar]
  • 261.Barbu AR, Bodin B, Welsh M, Jansson L, Welsh N. A perfusion protocol for highly efficient transduction of intact pancreatic islets of Langerhans. Diabetologia. 2006;49:2388–2391. doi: 10.1007/s00125-006-0390-5. [DOI] [PubMed] [Google Scholar]
  • 262.Kim SS, Ye C, Kumar P, Chiu I, Subramanya S, Wu H, Shankar P, Manjunath N. Targeted delivery of siRNA to macrophages for anti-inflammatory treatment. Mol Ther. 2010;18:993–1001. doi: 10.1038/mt.2010.27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Macedo GC, Magnani DM, Carvalho NB, Bruna-Romero O, Gazzinelli RT, Oliveira SC. Central role of MyD88-dependent dendritic cell maturation and proinflammatory cytokine production to control Brucella abortus infection. J Immunol. 2008;180:1080–1087. doi: 10.4049/jimmunol.180.2.1080. [DOI] [PubMed] [Google Scholar]

RESOURCES