Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Feb 6.
Published in final edited form as: Wiley Interdiscip Rev RNA. 2010 May 6;1(1):60–80. doi: 10.1002/wrna.1

Inflammation: Cytokines and RNA-based Regulation

Deborah J Stumpo 1, Wi S Lai 1, Perry J Blackshear 1,2
PMCID: PMC3915420  NIHMSID: NIHMS551239  PMID: 21956907

Abstract

The outcome of an inflammatory response depends upon the coordinated regulation of a variety of both pro-inflammatory and anti-inflammatory cytokines and other proteins. Regulation of these inflammation mediators can occur at multiple levels, including transcription, mRNA translation, post-translational modifications, and mRNA degradation. Post-transcriptional regulation has been shown to play an important role in controlling the expression of these mediators, allowing for normal initiation and resolution of the inflammatory response. Many inflammatory mediators have unstable mRNAs due, in part, to the presence of AU-rich elements in their 3′-untranslated regions. Increasing numbers of RNA-binding proteins have been identified that can bind to these AU-rich elements and then regulate the stability and/or translation of the mRNA. This review summarizes current knowledge about the role of several RNA-binding proteins that act through AU-rich elements to post-transcriptionally regulate the biosynthesis of proteins involved in inflammation.

Keywords: Cytokines, inflammation, AU-rich element, RNA-binding protein

Introduction

Inflammation is a normal defense mechanism that the body uses to protect itself from tissue injury. Regardless of the nature of the agent that caused the injury (physical, chemical, or microbial), the inflammatory response follows the same regulated process [1]. In the early or acute stage of inflammation, there is an initial increase in blood flow to the site of injury as well as increased vascular permeability. A variety of signaling molecules are released, including chemokines, cytokines, eicosanoids and adhesion molecules, resulting in the recruitment of polymorphonuclear leukocytes to the site of injury. As the inflammatory response progresses, monocytes, which differentiate locally into macrophages, and lymphocytes accumulate, assisting in the neutralization of the injurious agent. In the late stage of inflammation, tissue remodeling processes take over, leading to a resolution of the inflammatory response and a return to normal tissue physiology [2, 3, 4, 5, 6].

During acute inflammation, leukocytes that are recruited to the site of inflammation release pro-inflammatory mediators that initiate and enhance the acute phase of the response. The production of pro-inflammatory mediators (e.g., interleukin-1 [IL-1] and tumor necrosis factor alpha [TNF]) is counterbalanced by the local release of anti-inflammatory mediators such as interleukin-10 (IL-10) as well as endogenous anti-inflammatory molecules such as corticosteroids. The initiation and resolution of the inflammatory response involves the complex and coordinated expression of many factors, including cytokines, chemokines, growth factors, proteases, lipids, etc. Dysregulation of the normal regulatory pathways involved in the resolution of the inflammatory response can lead to the development of chronic inflammatory diseases such as rheumatoid arthritis and inflammatory bowel disease.

The production of both pro- and anti-inflammatory mediators of inflammation can be regulated at many levels, including gene transcription, mRNA translation, and mRNA degradation [7, 8, 9, 10, 11, 12]. Transcriptional activation is required to initiate the inflammatory response, but post-transcriptional mechanisms play important roles in controlling the expression of potentially harmful pro-inflammatory mediators.

In this review, we will discuss the post-transcriptional regulation of pro-inflammatory protein production, focusing on several mRNA-binding proteins that have been shown to participate in this regulation.

The role of AU-rich elements in mRNA stability

Many inflammatory mediators, such as cytokines and chemokines, have unstable mRNAs. This instability often results in part from the presence of cis-acting adenine and uridine-rich (AU-rich) elements, or AREs, in their 3′ untranslated (UTR) regions. These AU-rich elements provide binding sites for trans-acting factors, such as RNA-binding proteins, that can subsequently regulate the stability and/or translation of the mRNA. AREs were first identified as conserved regions within the 3′ UTRs of several human and mouse mRNAs encoding inflammatory-related proteins [13], including TNF, IL-1, granulocyte-macrophage colony-stimulating factor (GM-CSF), and interferon-alpha (IFN-α). Because 3′ UTRs are relatively poorly conserved compared to protein coding regions, it was proposed that the identified consensus sequence, UUAUUUAU, may play an important regulatory role. This regulatory function was demonstrated by Shaw and Kamen [14] when they showed that the placement of the ARE from the GM-CSF mRNA into the 3′ UTR of the relatively stable β-globin transcript caused the β-globin mRNA to become highly unstable.

The pro-inflammatory cytokine TNF has been known to play an important role in the pathogenesis of many human inflammatory diseases [15]. The use of anti-TNF agents has proven to be successful in the treatment of several inflammatory conditions including rheumatoid arthritis, Crohn’s disease, and psoriasis [16, 17], reinforcing the central role that TNF plays in chronic inflammation. TNF expression is regulated both at the transcriptional level as well as post-transcriptionally. Post-transcriptional control has been shown to occur both at the level of mRNA stability as well as translation [18]. The importance of the 3′ UTR of TNF in the development of arthritis was demonstrated in studies using mice that overexpressed a TNF/β-globin 3′ UTR transgene [19]. These mice expressed a human TNF gene driven by its own promoter and protein coding region, but the 3′ UTR from the human β-globin mRNA replaced the 3′ UTR of the human TNF transcript. These mice developed chronic inflammatory polyarthritis [19]. Treatment with an antibody to human TNF was able to prevent the development of disease, demonstrating the connection between TNF and the pathogenesis of the arthritis.

Subsequent animal studies established the role of the TNF 3′ UTR ARE in the development of inflammatory disease [20]. In these studies, the 69 bp ARE was deleted by homologous recombination “knockin” technology. Mice lacking this ARE (TNFΔARE mice) had increased constitutive and LPS-induced levels of circulating TNF protein and were more susceptible to the endotoxic effects of LPS [20]. Both peritoneal macrophages and bone marrow-derived macrophages (BMDM) expressed increased levels of TNF mRNA in the TNFΔARE mice compared to control mice although the level of TNF gene transcription was the same [20]. Actinomycin D studies showed that cells from the TNFΔARE mice accumulated more TNF mRNA than those from control mice, due to an increased TNF mRNA half-life [20]. Other studies have shown that the ARE in the 3′ UTR of TNF is also responsible for translational repression of TNF [20, 21, 22, 23]. This translational repression could be released upon treatment with lipopolysaccharide (LPS) [21]. Thus, the TNF ARE appeared to contribute to both TNF mRNA stability and translation. In the absence on the ARE, the TNFΔARE mice developed both chronic inflammatory arthritis and a Crohn’s-like inflammatory bowel disease, suggesting a link between excess TNF production and the analogous diseases in humans [20].

AREs within mRNAs are often between 50 and 150 nt in length and range greatly in composition. One classification was based on the number and distribution of AUUUA pentamers [24]. Class I AREs contain 1 to 3 AUUUA pentamer motifs scattered throughout the 3′ UTR within or near a U-rich region (e.g., c-myc and c-fos). Class II AREs contain at least two overlapping copies of the nonamer UAUUUA(U/A)(U/A) within a U-rich region (e.g., cytokine mRNAs such as TNF and GM-CSF). Class III AREs contain U-rich regions but no AUUUA pentamers (e.g., c-jun). Bakheet et al. [25] have created a database of human ARE mRNAs (ARED; http://brp.kfshrc.edu.sa/ARED/) based on the pattern of AUUUA pentamers. Group I contains 5 pentameric repeats, AUUUAUUUAUUUAUUUAUUUA; group II contains 4, AUUUAUUUAUUUAUUUA; group III contains 3 (A/U)AUUUAUUUAUUUA(A/U); group IV contains 2, (A/U)2AUUUAUUUA(A/U)2; and group V contains 1, (A/U)4AUUUA(A/U)4. Based on their analysis, the percentage of mRNAs within the human genome that contain AREs is about 8%. This database has been updated several times and the most recent update includes human, mouse and rat ARE mRNAs [26, 27, 28].

Pathways of ARE-mediated mRNA decay

Deadenylation, or the processive removal of the poly(A) tail, is thought to be the first and rate-limiting step in the degradation of mRNA in eukaryotes. Several eukaryotic deadenylases have been shown to participate in this process, including poly(A)-specific ribonuclease (PARN) and the CCR4/POP2/NOT1 complex (a complex of 9 proteins). AREs have been shown to enhance the rate of deadenylation [14, 29]. Following deadenylation, the mRNA “body” can be degraded further by decapping followed by 5′→3′ exonucleolytic decay; 3′→5′ exonucleolytic degradation; or by endonucleolytic degradation [12, 30, 31]. The 3′→5′ exonucleolytic decay pathway has been shown to require the exosome, a large multi-unit complex of 3′→5′ exonucleases [32, 33]. A component of the exosome, PM-SCL75, has been shown to interact directly with AREs [32]. In the decapping and 5′→3′ exonucleolytic decay pathway, the LSM 1–7 complex binds to the 3′ end of the mRNA and promotes decapping by the DCP1/DCP2 complex. Following decapping, the mRNA can be degraded by the 5′→3′ exonuclease XRN1 [12, 30, 31]. ARE-containing mRNAs have been shown to stimulate RNA decapping, suggesting that ARE-mediated decay can also occur through the 5′→3′ exonucleolytic decay pathway [34].

Several components of the 5′→3′ decay machinery, including deadenylases, the decapping enzymes DCP1/DCP2, the LSM 1–7 complex, XRN1 and others are enriched in granular cytoplasmic foci called processing bodies or P bodies [30, 31, 35, 36]. Translationally silent mRNA is also present in P bodies and mRNA decay intermediates have been detected in P bodies, leading to the proposal that P bodies are cellular sites of mRNA decay [37, 38]. P bodies form when mRNA translation is impaired or the 5′→3′ decay pathway gets overloaded. mRNA within P bodies can be degraded or held in a translationally inactive state until it is released to associate with polysomes for translation. Another type of RNA granule that can be found in eukaryotic cells is the stress granule or SG [39, 40]. Stress granules form in response to environmental stress (e.g., heat shock, oxidative stress, and UV irradiation) and contain stalled initiation complexes. Stress granule formation requires the phosphorylation of the initiation factor eIF2α. Phosphorylated eIF2α reduces the availability of the eIF2α-GTP-tRNAiMet ternary complex, resulting in blocked translation initiation and polysome disassembly [41, 42]. The components of SGs rapidly shuttle in and out of the SG [43]. Studies using live-cell microscopy demonstrated that P bodies and SGs can interact with each other [44]. It has been proposed that following polysome disassembly, released mRNA is sorted within SGs and some mRNAs are selected for storage or reinitiation while other mRNAs are selected to be delivered to P bodies where they are degraded [44]. A number of ARE-binding proteins have been shown (discussed below) to associate with components of the exosome and the 5′→3′ decay pathway as well as P bodies, suggesting that both degradation pathways are involved in ARE-mediated mRNA decay [33, 44, 45, 46, 47].

Post-transcriptional regulation via ARE-binding proteins

A growing number of ARE-binding proteins have been identified over the past 20 years [48]. Table I lists ARE-binding proteins that have been shown to regulate their target mRNAs at the post-transcriptional level. These proteins bind to AREs in their target RNAs via specific protein domains, including the RNA-recognition motif (RRM), the CCCH tandem zinc finger domain, and the K-homology domain (KH). The interactions between ARE-binding proteins and their target mRNAs have been shown using a variety of methods, including UV crosslinking, RNA electrophoretic mobility shift assay, co-transfection functional assays, RNA immunoprecipitation (RIP) assays, siRNA knockdown assays, or mRNA decay in cells isolated from ARE-binding protein knockout mice. To date, only relatively few ARE-binding proteins have been shown to be involved in inflammatory processes in intact animals.

Table I.

ARE-binding Proteins

ARE-binding protein RNA-binding domain Function References
APOBEC1 Unknown Destabilize mRNA [166, 167, 168, 169]
AUF1 (hnRNPD) RRM Destabilize mRNA
Stabilize mRNA
[115, 117, 118, 119, 140]
[121, 122]
AUF2 (CBF-A) RRM Stabilize mRNA [170]
CUGBP2 RRM Translational silencing
Stabilize mRNA
[171]
[172]
FXR1P K homology; RGG Translational silencing [173]
HuR (ELAVL1) RRM Stabilize mRNA
Translational activation
Translational silencing
[119, 135, 136, 137, 138, 139, 140, 142, 143, 144, 145, 146]
[119, 147, 148, 149]
[150]
KSRP K homology Destabilize mRNA [33, 95, 174, 175, 176, 177, 178]
NF90 DRBD Stabilize mRNA
Translational silencing
[179, 180, 181, 182, 183, 184]
[184]
RHAU Novel N-terminal RNA-binding domain Destabilize mRNA [185, 186]
YB1 CSD Stabilize mRNA [187, 188]
TIA-1 (TIA1) RRM Translational silencing [158, 159, 160, 161]
TIAR (TIAL1) RRM Translational silencing [157, 160, 162, 189, 190]
ZFP36 (TTP) CCCH zinc finger Destabilize mRNA
Stabilize mRNA
[59, 63, 64, 76, 77, 78, 79, 81, 82, 83, 84, 87, 88, 89, 92, 93, 96, 97, 99, 101, 102, 103, 191]
[94]
ZFP36L1 (BRF1) CCCH zinc finger Destabilize mRNA
Translational silencing
[63, 64, 82, 110, 111]
[108]
ZFP36L2 (BRF2) CCCH zinc finger Destabilize mRNA [51, 63, 64]
ZFP36L3 CCCH zinc finger Destabilize mRNA [52, 53]

APOBEC1, apolipoprotein B mRNA editing enzyme, catalytic polypeptide 1; AUF1/hnRNPD, AU-rich binding factor 1/heterogeneous nuclear ribonucleoprotein D; AUF-2/CBF-A, AU-rich binding factor 2/CArG-box-binding factor-A; CCCH, cysteine-cysteine-cysteine-histidine; CSD, cold shock domain; CUGBP2, CUG triplet repeat, RNA binding protein 2; DRBD, dsRNA-binding domain; FXR1P, fragile X-related protein 1; HuR/ELAVL1, Hu antigen R/embryonic lethal abnormal vision-like 1; KSRP, KH-type splicing regulatory protein; NF90, nuclear factor 90; RGG, arginine-glycine-glycine; RHAU, RNA helicase associated with AU-rich element; RRM, RNA recognition motif; YB1, Y-box-binding protein 1; TIA-1, T-cell restricted intracellular antigen 1; TIAR, TIA-1-related; ZFP36/TTP, zinc finger protein 36/tristetraprolin; ZFP36L1/BRF1, zinc finger protein 36-like 1/butyrate response factor 1; ZFP36L2/BRF2, zinc finger protein 36-like 2/butyrate response factor 2; ZFP36L3, zinc finger protein 36-like 3

The Tristetraprolin (TTP) family of proteins

Tristetraprolin (TTP), zinc finger protein 36 or ZFP36 (also known as Tis11, Nup475, and Gos24), is the prototype of a small family of four mammalian CCCH tandem zinc finger proteins capable of binding to RNA. TTP and two other family members, ZFP36L1 (TIS11B, cMG1, BRF1, ERF1, Berg36) and ZFP36L2 (TIS11D, BRF2, ERF2) are expressed in a variety of species, including man [49, 50, 51]. The most recently identified member, ZFP36L3, has only been found in rodents and has been detected only in placenta and extraembryonic tissues [52, 53]. The CCCH tandem zinc finger domain of all family member proteins is characterized by 2 zinc fingers, spaced 18 amino acids apart, each containing 3 cysteines and one histidine with a strictly defined internal spacing, C-X8-C-X5-C-X3-H [49]. TTP itself is the most extensively studied member of this family of proteins. It was first identified as the product of an immediate early response gene, Zfp36, in fibroblasts and other cells stimulated with insulin, phorbol esters, or serum [49, 54, 55, 56]. TTP is localized in the nucleus of serum-deprived quiescent fibroblasts and is rapidly translocated to the cytoplasm in response to mitogens or serum [57]. Translocation is accompanied by serine phosphorylation of the protein [58]. In macrophages, TTP is dramatically induced by LPS, and is then located primarily in the cytoplasm [59, 60].

Cell transfection and cell-free experiments have shown that TTP, as well as the three other mouse family members, ZFP36L1, ZFP36L2 and ZFP36L3, bind via their conserved tandem zinc finger domains to AREs in mRNAs at a consensus nonamer site, UUAUUUAUU. Binding of synthetic random zinc finger domain peptides to this sequence has been shown to occur with low nanomolar affinity [61]. This initial RNA binding leads to the deadenylation and subsequent degradation of the transcript [52, 53, 62, 63, 64, 65, 66, 67].

Several studies have implicated TTP in deadenylation as well as both 5′→3′ and 3′→5′ exonucleolytic decay. We have shown that TTP, as well as its other family members, can promote ARE-containing mRNA deadenylation in a cell-free assay [52, 65]. Co-transfection studies with PARN demonstrated that TTP’s ability to promote deadenylation was enhanced by adding increasing concentrations of PARN [65]. Chen et al. [33] demonstrated that the ARE-binding proteins AUF1 (AU-rich binding factor 1), KSRP (KH-type splicing regulatory protein) and TTP co-purified with human exosomes. Subsequent cell-free studies showed that the addition of KSRP or TTP to isolated exosomes stimulated ARE-mediated mRNA decay [33]. TTP and ZFP36L1 (BRF1) have been shown to interact with several components of the mRNA decay pathways including the deadenylase CCR4, the decapping enzymes DCP1 and DCP2, two components of the exosome, RRP4 and PM-SCL75, and the 5′→3′ exonuclease XRN1 [45, 46, 47]. These studies did not show an interaction between TTP or ZFP36L1 and PARN [45]. The N terminal domain of TTP and ZFP36L1was found to be the region responsible for interacting with the mRNA decay proteins [45], while both the N- and C-terminal domains were capable of activating mRNA decay when linked to a heterologous RNA-binding protein [45]. The N-terminal domain of TTP that was shown to interact with DCP2 was able to activate decapping of an ARE substrate in an in vitro decapping assay [46].

TTP and ZFP36L1 have also been shown to co-localize to P bodies along with components of the 5′→3′ mRNA decay pathway [44]. Subsequent studies demonstrated that TTP and ZFP36L1 were able to target a linked heterologous RNA-binding protein to P bodies [68]. The N- and C-terminal domains of TTP functioned as redundant P body localization domains [68]. These studies have lead to the hypothesis that TTP may exert its effect on ARE-mediated mRNA decay by recruiting the exosome to the bound transcript [33, 47]. TTP may also deliver bound transcripts to P bodies, where mRNA decay can proceed by the 5′→3′ pathway, or TTP may promote the assembly of P bodies through the recruitment of mRNA decay enzymes [9, 44, 68]. TTP has also been linked to the RNA-induced silencing complex (RISC). Jing et al. [69] have shown that TTP interacts with Argonaute/eiF2C2 and eiF2C4, components of the RISC complex. MiR16, a miRNA containing an UAAAUAUU sequence that is complimentary to the ARE in TNF mRNA, was required for ARE-mediated mRNA decay, and miR16 required the presence of TTP in order to have an effect on ARE-mediated decay [69]. It has been proposed that TTP, through binding to Argonaute proteins, assists miR16 targeting to the ARE, resulting in mRNA decay [69]. Argonaute proteins localize to P bodies and mRNAs that are targeted by miRNAs accumulate within P bodies [70].

The physiological role of TTP in intact animals was first suggested by studies involving TTP knockout (KO) mice. These mice appeared normal at birth but soon developed a complex syndrome that included inflammatory arthritis, cachexia, autoimmunity, dermatitis and myeloid hyperplasia [71]. Because some aspects of the TTP KO phenotype resembled mouse models of TNF overexpression [19, 72, 73], we treated newborn mice with a monoclonal antibody directed against mouse TNF [74]. This resulted in prevention of most aspects of the KO phenotype. Subsequent breeding experiments with mice deficient in both types of TNF receptor also alleviated most aspects of the KO phenotype [75]. Taken together, these studies indicated that TTP played a role in the production or clearance of TNF [75]. Subsequent studies done in macrophages isolated from TTP KO mice demonstrated that TTP could bind directly to the ARE within the TNF mRNA, and that TTP was capable of de-stabilizing the TNF mRNA [59]. These studies also showed that the tandem zinc finger domain of TTP was the region that interacted with the ARE of TNF. Thus, the increased production of both TNF protein and mRNA in the TTP KO mice was the result of an increase in the stability of TNF mRNA [59], resulting in overproduction of TNF protein.

Subsequent experiments identified a second physiological target of TTP, the GM-CSF mRNA, in bone marrow-derived stromal cells isolated from TTP KO mice. As was seen for TNF, the absence of TTP resulted in an increased production of GM-CSF due to a marked stabilization of GM-CSF mRNA [76]. Similar results were seen in stromal cells isolated from TTP KO mice that were also deficient in both TNF receptors, ruling out the possibility that increased circulating TNF was responsible for the observed effect on GM-CSF mRNA [76].

An important clue concerning TTP’s mechanism of action came out of these studies. Northern blot analysis of bone marrow-derived stromal cells demonstrated that, in cells from wild type (WT) mice, the GM-CSF mRNA exists as two species of similar abundance but differing in size by about 200 bases. This size difference was shown to be the result of the presence or absence of a poly(A) tail on the RNA [76]. The larger of the two species, the polyadenylated version, was almost the only transcript detectable in the cells from the TTP KO mice [76]. Actinomycin D mRNA turnover studies indicated that while both species of GM-CSF mRNA were unstable in wild type stromal cells, the fully polyadenylated transcript seen in the TTP KO stromal cells barely decayed at all during the course of the experiment [76]. Thus, the absence of TTP led to the failure of removal of the poly(A) tail (deadenylation), resulting in the accumulation of the fully polyadenylated GM-CSF mRNA and overproduction of the GM-CSF protein.

Many potential mRNA targets for TTP have since been described in the literature, using a variety of techniques (see Table II). For example, TTP overexpression studies have identified a number of mRNAs that were negatively regulated, and, in some cases, mRNA stability was shown to be decreased, including IL-3 [59, 77, 78, 79], plasminogen activator inhibitor type 2 or PAI-2 [80], cyclooxygenase-2 or Cox-2 [81], paired-like homeodomain transcription factor 2 or Pitx2 [82], IL-2 [78], IL-6 [78], IL-8 [83], vascular endothelial growth factor or VEGF [83], and coxsackie- and adenovirus-receptor-like membrane protein or CLMP [84]. A RIP-CHIP (immunoprecipitation with TTP antibody followed by RNA isolation and hybridization to Affymetrix GeneChips) approach was used to identify potential TTP targets in human dendritic cells [85]. More than 300 genes were identified by this approach of which 37 were present in the ARED2 database. Six of these identified targets were verified by real-time PCR and functional transfection studies. These targets include dual specificity phosphatase 1 (DUSP1, also known as mitogen-activated protein kinase phosphatase 1 or MKP-1), indoleamine 2,3-dioxygenase (IDO), superoxide dismutase 2 (SOD2), CD86, and MHC class I-B and F [85]. CD86 and MHC class I mRNAs were not in the ARED database, although CD86 contains two ARE-like sequences (UAUUUAU and UUAUUUUAU) in its 3′ UTR. Two previously identified targets of TTP, Cox-2 and DUSP1, were found to be transiently induced during the early stages of differentiation of 3T3-L1 preadipocytes [86]. Both of these transcripts contain AU-rich elements in the 3′ UTR and both were found to co-immunoprecipitate with TTP [86].

Table II.

TTP Family Members and Identified Target mRNAs

Protein mRNA Target Mechanism Binding Co-transfection Deadenylation siRNA knockdown KO cells References
TTP (ZFP36) TNFa mRNA stability REMSA; UVCL RNA; mRNA stability Cell-free RNA; Protein; mRNA stability [59, 62, 63, 64, 65, 66, 192]
GM-CSF mRNA stability REMSA; UVCL mRNA stability Cell-free; RNase H RNA; Protein; mRNA stability [64, 65, 76, 193]
IL-2 mRNA stability REMSA; UVCL mRNA stability RNA; Protein; mRNA stability [96]
Immediate early response gene 3 (Ier3) mRNA stability REMSA mRNA stability [101]
IL-10 mRNA stability RIP mRNA stability Cell-free RNA; Protein; mRNA stability [103, 191]
CXCL1 mRNA stability mRNA stability RNA; mRNA stability [99]
Polo-like kinase 3 (Plk3) mRNA stability REMSA mRNA stability [101, 102]
Interferon-α (IFN-α) mRNA stability REMSA; UVCL mRNA stability RNA; Protein; mRNA stability [97]
Cyclooxygenase -2 (Cox-2) mRNA stability UVCL; RIP RNA; Protein; mRNA stability [81, 191]
Myeloid/lymph oid or mixed lineage leukemia (trithorax) (Mllt11) mRNA stability mRNA stability [101]
Leukemia inhibitory factor (Lif) mRNA stability RIP mRNA stability mRNA stability [101, 103]
B double-prime 1 (Bdp1) mRNA stability mRNA stability [101]
Proviral integration site 3 (Pim3) mRNA stability mRNA stability [101]
RUN-and SH3 domain-containing 2 (Rusc2) mRNA stability mRNA stability [101]
Pleckstrin homology-like domain, family A, member 1 (Phlda1) mRNA stability mRNA stability [101]
IL-1α mRNA stability REMSA; UVCL RNA; mRNA stability [191]
IL-3 mRNA stability REMSA RNA; Protein; mRNA stability Rnase H [59, 64, 77, 78, 79, 193]
IL-6 mRNA stability RNA; Protein; mRNA stability Protein RNA; Protein; mRNA stability [78, 91, 191, 194]
Plasminogen activator inhibitor type 2 (PAI-2) mRNA stability UVCL RNA; Protein; mRNA stability [80]
Paired-like homeodomain transcription factor 2 (Pitx2) mRNA stability UVCL; RIP mRNA stabilityb [82]
1,4-galactosyltransferase mRNA stability UVCL; RIP mRNA stability RNA; mRNA stability [93]
Inducible nitric-oxide synthase (iNOS) mRNA stability UVCLc RNA; mRNA stability RNA; Protein [94]
Cyclin D1 mRNA stability REMSA mRNA stability [92]
Myc mRNA stability REMSA mRNA stability [92]
Cyclin-dependent kinase inhibitor 1A (P21) ND RIP RNA [90]
Fos ND RIP RNA RNA [90]
IL-12 ND Protein [91]
Macrophage inflammatory protein 2α (MIP2) ND Protein [91]
Chemokine (C-C) motif ligand 2 (CCL2) or monocyte chemotactic protein-1 (MCP1) ND RNA; Protein [194]
Chemokine (C-C) motif ligand 3 (CCL3) or macrophage inflammatory protein-1 (MIP1α) ND RNA; Protein [194]
Macrophage inflammatory protein-3α (MIP3α) ND Protein [91]
Dual specificity phosphatase 1 (DUSP1) ND RIP RNA [85]
VEGF mRNA stability REMSA; RIP; UVCL Protein; mRNA stability RNA [83, 88]
E2a-encoded transcription factor E47 (E47) ND REMSA RNA [89]
IL-8 mRNA stability UVCL Protein; mRNA stability Cell-free mRNA stability [83, 87]
Indolamine 2,3-dioxygenase (IDO) ND RIP RNA [85]
Superoxide dismutase (SOD2) ND RIP RNA [85]
CD86 ND RIP RNA [85]
MHC class I-B and F ND RIP RNA [85]
Coxsackie-and adenovirus-receptor-like membrane protein (CLMP) ND REMSA RNA [84]

ZFP36L1 VEGF translation REMSA mRNA stability mRNA stability RNA; Protein; mRNA stability; polysomes [108, 109]
TNF mRNA stability REMSA; UVCL RNA; mRNA stability Cell-free [63, 64, 65, 66]
GM-CSF mRNA stability mRNA stability Cell-free [64, 65]
IL-3 mRNA stability REMSA mRNA stability Cell-free RNA [64, 110]
Pitx2 mRNA stability UVCL; RIP mRNA stabilityb [82]
Human inhibitor of apoptosis protein-2 (c-IAP2) ND RNA [195]
Steroidogenic acute regulatory protein (STAR) mRNA stability UVCL; RIP RNA RNA; Protein [111]

ZFP36L2 TNF mRNA stability REMSA; UVCL RNA; mRNA stability Cell-free [63, 64, 65, 66]
GM-CSF mRNA stability mRNA stability [64]
IL-3 mRNA stability mRNA stability [64]
CXCL1 ND RNA [51]

ZFP36L3 TNF mRNA stability REMSA mRNA stability Cell-free [52, 53]
a

Confirmed “physiological” targets are listed in bold; mRNA stability has been demonstrated in cells derived from knockout animals

b

In vitro RNA decay system

c

No direct binding observed between TTP and iNOS mRNA

ND, Not determined

Cell-free, cell-free deadenylation assay; KO cells, assays performed in cells isolated from knockout animals; Polysomes, mRNA association with polysomes; Protein, protein measured by western or ELISA; REMSA, RNA electrophoretic mobility shift assay; RIP, RNA immunoprecipitation; RNA; RNA measured by northern, RT-PCR, or RNase protection; RNase H; removal of poly(A) tail measured by hybridization to oligo(dT) followed by RNase H digestion; UVCL, UV cross linking assay

Knockdown of TTP by siRNA has been performed in a variety of cell types, permitting the identification of mRNAs whose stability is regulated by TTP. These include IL-8 [87], VEGF [88], E2a-encoded transcription factor E47 [89], c-fos [90], cyclin-dependent kinase inhibitor 1A (p21; [90]), IL-6 [91], IL-12 [91], Cxcl2 or macrophage inflammatory protein 2-alpha (MIP-2; [91]), cyclin D1 [92], c-myc [92], and 1,4-galactosyltransferase [93]. All of these transcripts were found to be negatively affected by TTP, i.e., TTP caused the destabilization of these mRNAs. In some cases mRNA (p21, c-fos) or protein (IL-12, MIP-2) levels were measured but mRNA stability was not assessed [90, 91].

In contrast to the general theme of TTP as an mRNA destabilizing factor, two mRNAs have been found to be stabilized by TTP using siRNA knockdown approaches [91, 94]. In one case, involving inducible nitric-oxide synthase (iNOS) mRNA [94], TTP did not interact with the iNOS mRNA itself (based on UV crosslinking), but instead interacted with the KH-type splicing regulatory protein, KSRP [94], which has been shown to destabilize iNOS mRNA through binding to the ARE in the 3′UTR of the transcript [95]. Knockdown of TTP in a mouse macrophage cell line resulted in decreased production of MIP-3α, also known as CC chemokine ligand 20 (CCL20) [91]. MIP-3α does not contain classical nonamer binding sites for TTP, but 5 AUUUA pentamers are present in the 3′ UTR. Binding studies were not performed in this study nor were mRNA levels or stability measured [91]. These studies suggest that TTP may sometimes regulate mRNAs by indirect mechanisms, for example, involving protein-protein interactions.

Since the identification of TNF and GM-CSF as the first documented “physiological” targets of TTP, a number of target mRNAs have been validated as physiologically relevant, i.e., meeting the strict criterion that they are stabilized in cells from TTP KO mice (see Table II, bold entries).

For example, Ogilvie et al. [96] examined the role of TTP in the regulation of IL-2 mRNA in T cells isolated from TTP KO mice. First, they demonstrated that in activated human T cells, TTP is induced and can bind to the ARE in the 3′ UTR of IL-2 [96]. Activated total splenocytes or CD-3 positive T cells from TTP KO mice produced more IL-2 protein and mRNA compared to WT cells [96]. Subsequent actinomycin D decay assays demonstrated that lack of TTP resulted in stabilization of IL-2 mRNA [96]. In a subsequent study, Ogilvie et al. [97] showed that activated T cells purified from TTP KO spleens produced more interferon-γ (IFN-γ) than WT cells, again associated with increased mRNA stability seen in the absence of TTP [97]. Both exogenous and endogenous TTP could bind to a 70 nucleotide ARE present in the 3′ UTR of IFN-γ [97].

Cxcl1 (also known as KC, growth-related oncogene α, Gro-α, and melanoma growth stimulatory activity) is a member of the CXC subfamily of chemokines and acts as a functional homologue of human interleukin-8 in the mouse [98]. The 3′ UTR of Cxcl1 contains a cluster of several overlapping AUUUA pentamers as well as isolated pentamers. These AREs confer instability to the Cxcl1 mRNA [99, 100]. Co-transfection studies showed that TTP could interact with the AREs in the 3′ UTR of the Cxcl1 mRNA, resulting in a decreased mRNA half-life [99]. Cxcl1 mRNA was found to be stabilized in peritoneal macrophages isolated from TTP KO mice [99].

Two different “global” approaches aimed at identifying new targets of TTP have recently been published. In one approach, microarray analysis of RNA isolated from WT and TTP KO fibroblasts was used to identify transcripts that had different decay rates following serum stimulation and actinomycin D treatment [101, 102, 103]. Lai et al. [101] identified 250 transcripts apparently stabilized in the TTP knockout fibroblasts. Of those, 23 had two or more UAUUUAU binding sites that were conserved in the human transcript, and nine of those were confirmed to be stabilized by northern blot analysis. One of the nine, Ier3, is a protein that has been implicated in the physiological control of blood pressure [104]. Another is Plk3, a serine-threonine kinase that has been implicated in many cellular processes, including cell cycle progression, mitosis, cytokinesis and the response to DNA damage [105, 106, 107]. Actinomycin D mRNA decay assays done in WT and TTP KO fibroblasts verified that Ier3 and Plk3 mRNAs were stabilized in the absence of TTP [101, 102]. One of the TTP family members, ZFP36L1, was found not to have an effect on either Ier3 or Plk3 mRNA stability [101, 102]. As expected, the AREs from Ier3 and Plk3 could bind TTP, and could mediate TTP-dependent mRNA decay in co-transfection experiments [101, 102].

In a second “global” experiment, microarray analysis was used to identify transcripts that could be co-immunoprecipitated with TTP in LPS treated RAW264.7 macrophages [103]. This approach identified a number of potential TTP-associated mRNAs, of which only a subset contained AUUUA pentamers or UUAUUUAUU nonamers. One of these, IL-10 was identified as a physiological target of TTP using BMDM and splenocytes from WT and TTP KO mice [103]. Actinomycin D studies demonstrated that lack of TTP decreased the rate of IL-10 mRNA decay [103]. In both “global” experiments, many transcripts were identified that did not contain an obvious ARE, raising the possibility that TTP could regulate some transcripts indirectly.

These studies have demonstrated that TTP can regulate both pro-inflammatory and anti-inflammatory mRNA decay, which may give TTP a way to “fine tune” the inflammatory response and prevent the development of chronic inflammation. It will be important to understand the timing of these countervailing responses in the acute response to an inflammatory stimulus, and the return to homeostasis.

As mentioned earlier, mammalian members of the TTP family of proteins behave like TTP in cell-free RNA binding assays, co-transfection assays evaluating the turnover of ARE-containing reporter transcripts, and cell-free deadenylation assays [52, 53, 62, 63, 64, 65, 66, 67]. Studies using TTP KO fibroblasts have demonstrated at least two examples of known TTP target transcripts whose stability was not affected by ZFP36L1 (Ier3 and Plk3 [101, 102]) or ZFP36L2 (Plk3 [102]), at least in this experimental system. Although there may be overlapping biochemical functions among the different TTP family members, their different physiological roles in the intact mouse probably depend on differences in regulated and tissue-specific expression, as well as probable differences in post-translational modifications, binding protein association, subcellular localization, etc.

The different physiological roles are illustrated by the different phenotypes of the Zfp36l1 [50, 108] and Zfp36l2 knockout mice [51]. Zfp36l1 [50] knockout embryos died between embryonic days 8 and 10, primarily due to failure of chorioallantoic fusion [50]. When fusion did occur, the KO placentas exhibited decreased cell proliferation and atrophy of the trophoblast layers [50]. No physiological target transcripts were identified in this study, but we postulated that ZFP36L1 may regulate one or more mRNAs whose encoded proteins may be involved in placental development. Bell et al. [108] also saw universal embryonic lethality at mid-gestation, with trophoblast layer defects in the placenta. These authors identified VEGF-A as a target mRNA for ZFP36L1 [108]. VEGF-A contains a prominent ARE in its 3′ UTR; previous studies had shown that ZFP36L1 could bind to this ARE, and that overexpression of ZFP36L1 caused decreases stability of VEGF-A mRNA [109]. Subsequent experiments demonstrated that there was no change in the stability of VEGF-A mRNA but rather there was enhanced association with polysomes, suggesting that ZFP36L1 regulated VEGF-A at the level of translation [108]. It appears from both reports that the determination of the role of ZFP36L1 in the adult animal will require the use of conditional KO mice.

Overexpression of ZFP36L1, like TTP, can result in increased turnover of TNF, GM-CSF and IL-3 mRNA [63, 64, 110]. More recently, ZFP36L1 has been shown to regulate the transcript for steroidogenic acute regulatory (STAR) protein. ZFP36L1 interacted with the ARE nonamer repeats present in the 3′ UTR, and ZFP36L1 overexpression and siRNA knockdown resulted in decreased and increased levels of STAR mRNA, respectively [111]. Interestingly, the effect of ZFP36L1 overexpression and knockdown had the opposite effect on protein levels, i.e., overexpression increased STAR protein levels and knockdown decreased protein levels [111]. The authors suggested that ZFP36L1 was capable of decreasing the stability of STAR mRNA while simultaneously enhancing translation, leading to increased cholesterol metabolism [111].

A partial Zfp36l2 KO mouse was described, in which the amino terminal 29 amino acids were deleted [112]. These mutant mice, ΔN-ZFP36L2, expressed decreased levels of an amino-terminal truncated protein that still contained an intact, functional tandem zinc finger domain. These mice displayed complete female infertility due to a defect in early embryonic development [112], presumably because of a defect in stored maternal mRNA and protein in the fertilized egg and early embryo. We recently described a full Zfp36l2 KO mouse with a very different phenotype [51]. Mice completely lacking ZFP36L2 exhibited profound and lethal hematopoietic defects. There were decreased definitive hematopoietic progenitors in fetal livers and yolk sac, and fetal liver hematopoietic stem cells were unable to reconstitute the hematopoietic system of lethally irradiated recipients [51]. Microarray analysis of RNA from fetal livers identified the transcript for the chemokine Cxcl1, among others, as being upregulated in the KO mice [51]. Actinomycin D mRNA decay assays showed no apparent change in the turnover of Cxcl1 mRNA in KO mouse fibroblasts; however, fibroblasts may not be the appropriate cell source in which to study Cxcl1 mRNA stability [51]. In addition, it may be that translation rather than mRNA stability is affected, as suggested for VEGF-A in Zfp36l1 KO fibroblasts [108]. Future studies to elucidate ZFP36L2 target transcripts will involve analysis of gene expression in fetal liver hematopoietic stem cells as well as conditional KO strategies. It should be noted that overexpression of Zfp36L2, like TTP and ZFP36L1, can promote the turnover of TNF, GM-CSF and IL-3 mRNA [63, 64], and that ZFP36L2, like TTP and ZFP36L1, can bind to 14-3-3 proteins [113] and can co-localize in P bodies with decapping enzymes [36].

AUF1

AUF1 (ARE/poly-(U) binding degradation factor 1 or heterogeneous nuclear ribonucleoprotein D, HNRNPD) was the first identified ARE-binding protein [114, 115, 116]. AUF1 has been shown to have both destabilizing effects on mRNA based on overexpression and knockdown experiments (transcripts of c-myc [115]), c-fos and GM-CSF [117], IL-3 [118], p21 and cyclin D1 [119], and iNOS [120]), as well as stabilizing effects based on overexpression studies (c-myc, c-fos, GM-CSF, TNF [121], parathyroid hormone [122]). AUF1 exists in 4 different isoforms (p37, p40, p42 and p45) resulting from alternative splicing, and each isoform apparently shows a different affinity for ARE-containing mRNAs, p37>p42>p45>p40 [123]. All four isoforms contain two nonidentical RNA recognition motifs, RRM1 and RRM2 [123]. There is differential expression of the different isoforms in some cell types, and it has been hypothesized that the opposing effects of AUF1 on mRNA stability may result from the relative levels of expression of each isoform in a given cell type or in response to a given stimulus [118, 124, 125]. The p37 isoform has been shown to interact with the exosome [33] and to exhibit the greatest destabilizing activity towards ARE-containing mRNAs [126]. Like the TTP family members, AUF1 can shuttle between the nucleus and cytoplasm [116, 119, 127, 128]. AUF1 KO mice are more sensitive to LPS-induced endotoxemia than WT mice [129], and serum levels of the pro-inflammatory cytokines, TNF and IL-1β, were found to be increased in the KO mice. Studies in LPS-stimulated peritoneal macrophages demonstrated that this increased production of TNF and IL-1β was the result of increased mRNA stability [129]. Treatment with neutralizing antibodies to both TNF and IL-1β together prior to the LPS challenge was able to protect the KO mice from lethal endotoxic shock [129]. These studies showed that AUF1 exerted post-transcriptional control over pro-inflammatory cytokine expression. A more recent report demonstrated that the AUF1 KO mice developed chronic pruritic dermatitis with age that resembles some aspects of human atopic dermatitis [130]. AUF1 KO mice also exhibited increased contact hypersensitivity characterized by increased T cell and macrophage infiltration and an increase in the production of pro-inflammatory cytokines TNF, IL-2 and IL-1β, and the monocyte chemoattractant protein 1 (MCP1; CCL2) [130]). These studies link AUF1 and its regulation of pro-inflammatory mediator mRNAs with the inflammatory response seen in skin.

Hur

Hur (Hu antigen R, HuA, ELAVL1) is the only ubiquitously expressed member of the mammalian embryonic lethal abnormal vision (ELAV) family of RNA-binding proteins [131, 132]. Expression of the other 3 members of this family, HuB (Hel-N1, ELAVL2), HuC (ELAVL3), and HuD (ELAVL4), is primarily restricted to neuronal cells [133]. HuR contains 3 RRMs and, like AUF1, is primarily nuclear but can shuttle between the nucleus and cytoplasm [131, 134, 135]. Numerous studies involving overexpression and RNA knockdown of HuR have implicated HuR in the stabilization many mRNAs, including TNF [136], VEGF [137, 138], IL-8 [138], Cox-2 [139], GM-CSF, c-fos [140, 141], p21 [119, 142], Cyclin A [143], Cyclin B1 [143], Cyclin D1 [119], iNOS [144], activating transcription factor-2 or ATF-2 [145], and the X chromosome-linked inhibitor of apoptosis protein, XIAP [146]. HuR has also been implicated in enhancing the translation of several mRNAs, including p21 [119], p53 [147, 148] and hypoxia-inducible factor 1α or HIF-1α [149]. Overexpression of HuR in vivo resulted in translational silencing of some select pro-inflammatory mediators. Katsanou et al. [150] established conditional myeloid-specific overexpressing HuR transgenic mice. They demonstrated in LPS-stimulated peritoneal macrophages that the myeloid-specific expression of HuR resulted in increased mRNA stability for TNF and Cox-2 but actually inhibited translation of the stabilized mRNA [150]. They also saw translational silencing of transforming growth factor β1 (TGFβ1) mRNA [150]. Administration of concanavalin A to control mice resulted in inflammatory hepatitis and liver damage due to the release of inflammatory mediators; similar treatment of mice overexpressing HuR did not result in inflammatory hepatitis or liver damage [150]. These studies suggest a link between HuR’s translational silencing of pro-inflammatory mediators and the suppression of inflammatory responses in the mouse.

Two descriptions of HuR KO mice were recently published. In the first report, Katsanou et al. [151] demonstrated that HuR was essential for normal mouse embryonic development, and that HuR KO mice died at mid-gestation due to placental defects. A second KO has been described in which HuR deficiency was also embryonic lethal [152]. The postnatal role of HuR was investigated using a tamoxifen-inducible Cre-mediated gene deletion strategy. Postnatal deletion of HuR resulted in atrophy of hematopoietic and lymphoid tissues, accompanied by decreased cellularity as a result of increased apoptosis in progenitor populations [152]. KO animals developed cachexia and died within 10 days of treatment with tamoxifen. Subsequent studies with embryonic fibroblasts demonstrated that, in the absence of HuR, p53 levels were increased due to decreased levels of the p53 upstream regulator, Mdm2 [152]. HuR binds to the 3′ UTR ARE in Mdm2, resulting in stabilization of the mRNA; thus, in HuR KO fibroblasts, Mdm2 mRNA is destabilized [152]. HuR is highly expressed in tissues undergoing active proliferation and the authors suggest that HuR may play an important survival role in actively dividing cell populations. Studies done so far with this conditional KO have not identified a connection between HuR and inflammatory processes, but tissue- or cell-specific KO of HuR, as opposed to a global KO, may shed more light on the post-transcriptional action of HuR on pro-inflammatory mediators.

Translational regulation by ARE-binding proteins

TIA-1/TIAR

The T-cell restricted intracellular antigen 1 (TIA-1 or TIA1) and the TIA-1-related protein (TIAR or TIAL1) are closely related RNA-binding proteins containing 3 RRMs that have been implicated in the apoptotic process as well as RNA binding [153, 154, 155, 156]. TIA-1 and TIAR have been shown to repress the translation of both pro-inflammatory [157, 158, 159, 160, 161] as well as anti-inflammatory mRNAs [162]. The physiological relevance of TIA-1 translational silencing of target mRNAs was demonstrated in TIA-1 KO mice. Mice deficient in TIA-1 were phenotypically normal in a BALB/c background [158]. LPS-stimulated peritoneal macrophages from TIA-1 KO mice expressed more TNF protein than did WT macrophages, although there was no difference in the levels or stability of TNF mRNA [158]. The absence of TIA-1 resulted in an increase in the proportion of TNF transcripts associated with polysomes, suggesting that TIA-1 functions as a translational silencer [158]. The translational silencing effect of TIA-1 appeared to be relatively selective for TNF, as no effect was seen for GM-CSF, IL-1β, or IFN-γ [158]. TIA-1 KO mice were more sensitive to endotoxic shock induced by LPS than WT mice [158].

The TIA-1 KO phenotype was different in a C57Bl/6 background. Macrophages from TIA-1 KO mice in a C57Bl/6 background produced significantly more TNF protein in response to LPS than was seen in a BALB/c background [161, 163]. It was observed that TIA-1 had no translational silencing effect on TNF in activated T cells suggesting that TIA-1 exerts its post-transcriptional effect in a cell-specific manner [163]. C57Bl/6 TIA-1 KO mice developed mild arthritis, which was not seen in a BALB/c background [161]. TIA-1/TTP double KO mice developed a more severe arthritis than was seen in either the TIA-1 or TTP KO mice, suggesting that both proteins have an influence on the severity of arthritis that develops [161]. Interestingly, bone marrow neutrophils and not macrophages were the main source of the increased production of TNF protein seen in the double KO mice [161]. These studies showed that TIA-1 and TTP could act as genetic modifiers of inflammatory arthritis that, through different post-transcriptional mechanisms, mRNA stability and mRNA translation, can regulate the abundance of pro-inflammatory mediators.

In addition to selectively interacting with ARE-containing mRNAs, TIA-1 and TIAR are thought to be involved in the general repression of translation seen during the stress response [164]. TIA-1 and TIAR shuttle between the nucleus and cytoplasm, and under stress conditions they can promote the assembly of translationally inactive preinitiation complexes and the formation of stress granules within the cytoplasm [41]. The mechanism by which TIA-1 and TIAR regulate the translation of selective ARE-containing mRNAs is thought to occur by the same mechanism as used for general translational repression in response to stress [165].

Conclusions

Post-transcriptional regulation of the many pro- and anti-inflammatory proteins involved in the initiation and subsequent resolution of the inflammatory response can influence the severity of the response. A major determining factor in the post-transcriptional control of many cytokines is the presence of a 3′ UTR AU-rich region in their transcripts. Both mRNA stability/instability and translational suppression can be mediated by ARE-binding proteins, several of which have been discussed in this review. Many ARE-binding proteins have been identified since the cloning of the first ARE-binding protein, AUF1 (see Table I, [48]). In vitro assays have identified target mRNAs for many ARE-binding proteins and also revealed many aspects of the mechanisms by which ARE-binding proteins regulate their target mRNAs. Studies in intact animals have begun to elucidate the roles of specific ARE-binding proteins and their target mRNAs in physiological and pathological inflammation. To our knowledge, TTP was the first ARE-binding protein demonstrated to have an effect on inflammation in intact animals, due to its effect on the regulation of TNF mRNA stability [59, 71]. Since then, several other ARE-binding proteins have been shown to be involved in the inflammatory response in vivo [130, 150, 161]. Due to the deleterious effect of the complete absence of some of the ARE-binding proteins on embryonic development, the demonstration of their possible involvement in inflammation and the identification of physiologically relevant target mRNAs in the adult will have to wait for conditional cell- or tissue-specific KO targeting approaches.

Given the large number of ARE-binding proteins, and the number of target mRNAs that can be regulated by multiple ARE-binding proteins, it seems likely that any given mRNA will be regulated by multiple ARE-binding proteins in intact animals. It will be fascinating to determine how these multiple ARE-binding proteins interact with AREs and other ARE-binding proteins to bring about the degree of gene regulation needed to respond to an inflammatory stimulus.

Acknowledgments

Supported in part by the Intramural Research Program of the NIH, NIEHS.

References

  • 1.Ryan GB, Majno G. Acute inflammation. A review. Am J Pathol. 1977;86:183–276. [PMC free article] [PubMed] [Google Scholar]
  • 2.Serhan CN, Brain SD, Buckley CD, Gilroy DW, Haslett C, O’Neill LA, Perretti M, Rossi AG, Wallace JL. Resolution of inflammation: state of the art, definitions and terms. Faseb J. 2007;21:325–332. doi: 10.1096/fj.06-7227rev. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Chin AC, Parkos CA. Pathobiology of neutrophil transepithelial migration: implications in mediating epithelial injury. Annu Rev Pathol. 2007;2:111–143. doi: 10.1146/annurev.pathol.2.010506.091944. [DOI] [PubMed] [Google Scholar]
  • 4.Calder PC, Albers R, Antoine JM, Blum S, Bourdet-Sicard R, Ferns GA, Folkerts G, Friedmann PS, Frost GS, Guarner F, Lovik M, Macfarlane S, Meyer PD, M’Rabet L, Serafini M, van Eden W, van Loo J, Vas Dias W, Vidry S, Winklhofer-Roob BM, Zhao J. Inflammatory disease processes and interactions with nutrition. Br J Nutr. 2009;101 (Suppl 1):S1–45. doi: 10.1017/S0007114509377867. [DOI] [PubMed] [Google Scholar]
  • 5.Stramer BM, Mori R, Martin P. The inflammation-fibrosis link? A Jekyll and Hyde role for blood cells during wound repair. J Invest Dermatol. 2007;127:1009–1017. doi: 10.1038/sj.jid.5700811. [DOI] [PubMed] [Google Scholar]
  • 6.Hao S, Baltimore D. The stability of mRNA influences the temporal order of the induction of genes encoding inflammatory molecules. Nat Immunol. 2009;10:281–288. doi: 10.1038/ni.1699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Stoecklin G, Anderson P. Posttranscriptional mechanisms regulating the inflammatory response. Adv Immunol. 2006;89:1–37. doi: 10.1016/S0065-2776(05)89001-7. [DOI] [PubMed] [Google Scholar]
  • 8.Katsanou V, Dimitriou M, Kontoyiannis DL. Post-transcriptional regulators in inflammation: exploring new avenues in biological therapeutics. Ernst Schering Found Symp Proc. 2006:37–57. doi: 10.1007/2789_2007_038. [DOI] [PubMed] [Google Scholar]
  • 9.Anderson P. Post-transcriptional control of cytokine production. Nat Immunol. 2008;9:353–359. doi: 10.1038/ni1584. [DOI] [PubMed] [Google Scholar]
  • 10.Zhang T, Kruys V, Huez G, Gueydan C. AU-rich element-mediated translational control: complexity and multiple activities of trans-activating factors. Biochem Soc Trans. 2002;30:952–958. doi: 10.1042/bst0300952. [DOI] [PubMed] [Google Scholar]
  • 11.Saklatvala J, Dean J, Clark A. Control of the expression of inflammatory response genes. Biochem Soc Symp. 2003:95–106. doi: 10.1042/bss0700095. [DOI] [PubMed] [Google Scholar]
  • 12.Hollams EM, Giles KM, Thomson AM, Leedman PJ. MRNA stability and the control of gene expression: implications for human disease. Neurochem Res. 2002;27:957–980. doi: 10.1023/a:1020992418511. [DOI] [PubMed] [Google Scholar]
  • 13.Caput D, Beutler B, Hartog K, Thayer R, Brown-Shimer S, Cerami A. Identification of a common nucleotide sequence in the 3′-untranslated region of mRNA molecules specifying inflammatory mediators. Proc Natl Acad Sci U S A. 1986;83:1670–1674. doi: 10.1073/pnas.83.6.1670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Shaw G, Kamen R. A conserved AU sequence from the 3′ untranslated region of GM-CSF mRNA mediates selective mRNA degradation. Cell. 1986;46:659–667. doi: 10.1016/0092-8674(86)90341-7. [DOI] [PubMed] [Google Scholar]
  • 15.Vassalli P. The pathophysiology of tumor necrosis factors. Annu Rev Immunol. 1992;10:411–452. doi: 10.1146/annurev.iy.10.040192.002211. [DOI] [PubMed] [Google Scholar]
  • 16.Taylor PC, Feldmann M. Anti-TNF biologic agents: still the therapy of choice for rheumatoid arthritis. Nat Rev Rheumatol. 2009;5:578–582. doi: 10.1038/nrrheum.2009.181. [DOI] [PubMed] [Google Scholar]
  • 17.Bradley JR. TNF-mediated inflammatory disease. J Pathol. 2008;214:149–160. doi: 10.1002/path.2287. [DOI] [PubMed] [Google Scholar]
  • 18.Beutler B, Cerami A. The biology of cachectin/TNF--a primary mediator of the host response. Annu Rev Immunol. 1989;7:625–655. doi: 10.1146/annurev.iy.07.040189.003205. [DOI] [PubMed] [Google Scholar]
  • 19.Keffer J, Probert L, Cazlaris H, Georgopoulos S, Kaslaris E, Kioussis D, Kollias G. Transgenic mice expressing human tumour necrosis factor: a predictive genetic model of arthritis. Embo J. 1991;10:4025–4031. doi: 10.1002/j.1460-2075.1991.tb04978.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Kontoyiannis D, Pasparakis M, Pizarro TT, Cominelli F, Kollias G. Impaired on/off regulation of TNF biosynthesis in mice lacking TNF AU-rich elements: implications for joint and gut-associated immunopathologies. Immunity. 1999;10:387–398. doi: 10.1016/s1074-7613(00)80038-2. [DOI] [PubMed] [Google Scholar]
  • 21.Han J, Brown T, Beutler B. Endotoxin-responsive sequences control cachectin/tumor necrosis factor biosynthesis at the translational level. J Exp Med. 1990;171:465–475. doi: 10.1084/jem.171.2.465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Kruys V, Kemmer K, Shakhov A, Jongeneel V, Beutler B. Constitutive activity of the tumor necrosis factor promoter is canceled by the 3′ untranslated region in nonmacrophage cell lines; a trans-dominant factor overcomes this suppressive effect. Proc Natl Acad Sci U S A. 1992;89:673–677. doi: 10.1073/pnas.89.2.673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Kruys V, Huez G. Translational control of cytokine expression by 3′ UA-rich sequences. Biochimie. 1994;76:862–866. doi: 10.1016/0300-9084(94)90188-0. [DOI] [PubMed] [Google Scholar]
  • 24.Chen CY, Shyu AB. AU-rich elements: characterization and importance in mRNA degradation. Trends Biochem Sci. 1995;20:465–470. doi: 10.1016/s0968-0004(00)89102-1. [DOI] [PubMed] [Google Scholar]
  • 25.Bakheet T, Frevel M, Williams BR, Greer W, Khabar KS. ARED: human AU-rich element-containing mRNA database reveals an unexpectedly diverse functional repertoire of encoded proteins. Nucleic Acids Res. 2001;29:246–254. doi: 10.1093/nar/29.1.246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Bakheet T, Williams BR, Khabar KS. ARED 2.0: an update of AU-rich element mRNA database. Nucleic Acids Res. 2003;31:421–423. doi: 10.1093/nar/gkg023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Bakheet T, Williams BR, Khabar KS. ARED 3.0: the large and diverse AU-rich transcriptome. Nucleic Acids Res. 2006;34:D111–114. doi: 10.1093/nar/gkj052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Halees AS, El-Badrawi R, Khabar KS. ARED Organism: expansion of ARED reveals AU-rich element cluster variations between human and mouse. Nucleic Acids Res. 2008;36:D137–140. doi: 10.1093/nar/gkm959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Ford LP, Wilusz J. An in vitro system using HeLa cytoplasmic extracts that reproduces regulated mRNA stability. Methods. 1999;17:21–27. doi: 10.1006/meth.1998.0703. [DOI] [PubMed] [Google Scholar]
  • 30.Parker R, Sheth U. P bodies and the control of mRNA translation and degradation. Mol Cell. 2007;25:635–646. doi: 10.1016/j.molcel.2007.02.011. [DOI] [PubMed] [Google Scholar]
  • 31.Garneau NL, Wilusz J, Wilusz CJ. The highways and byways of mRNA decay. Nat Rev Mol Cell Biol. 2007;8:113–126. doi: 10.1038/nrm2104. [DOI] [PubMed] [Google Scholar]
  • 32.Mukherjee D, Gao M, O’Connor JP, Raijmakers R, Pruijn G, Lutz CS, Wilusz J. The mammalian exosome mediates the efficient degradation of mRNAs that contain AU-rich elements. Embo J. 2002;21:165–174. doi: 10.1093/emboj/21.1.165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Chen CY, Gherzi R, Ong SE, Chan EL, Raijmakers R, Pruijn GJ, Stoecklin G, Moroni C, Mann M, Karin M. AU binding proteins recruit the exosome to degrade ARE-containing mRNAs. Cell. 2001;107:451–464. doi: 10.1016/s0092-8674(01)00578-5. [DOI] [PubMed] [Google Scholar]
  • 34.Gao M, Wilusz CJ, Peltz SW, Wilusz J. A novel mRNA-decapping activity in HeLa cytoplasmic extracts is regulated by AU-rich elements. Embo J. 2001;20:1134–1143. doi: 10.1093/emboj/20.5.1134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Anderson P, Kedersha N. Stress granules. Curr Biol. 2009;19:R397–398. doi: 10.1016/j.cub.2009.03.013. [DOI] [PubMed] [Google Scholar]
  • 36.Stoecklin G, Anderson P. In a tight spot: ARE-mRNAs at processing bodies. Genes Dev. 2007;21:627–631. doi: 10.1101/gad.1538807. [DOI] [PubMed] [Google Scholar]
  • 37.Sheth U, Parker R. Decapping and decay of messenger RNA occur in cytoplasmic processing bodies. Science. 2003;300:805–808. doi: 10.1126/science.1082320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Cougot N, Babajko S, Seraphin B. Cytoplasmic foci are sites of mRNA decay in human cells. J Cell Biol. 2004;165:31–40. doi: 10.1083/jcb.200309008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Anderson P, Kedersha N. Stress granules: the Tao of RNA triage. Trends Biochem Sci. 2008;33:141–150. doi: 10.1016/j.tibs.2007.12.003. [DOI] [PubMed] [Google Scholar]
  • 40.Anderson P, Kedersha N. RNA granules: post-transcriptional and epigenetic modulators of gene expression. Nat Rev Mol Cell Biol. 2009;10:430–436. doi: 10.1038/nrm2694. [DOI] [PubMed] [Google Scholar]
  • 41.Kedersha NL, Gupta M, Li W, Miller I, Anderson P. RNA-binding proteins TIA-1 and TIAR link the phosphorylation of eIF-2 alpha to the assembly of mammalian stress granules. J Cell Biol. 1999;147:1431–1442. doi: 10.1083/jcb.147.7.1431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Anderson P, Kedersha N. RNA granules. J Cell Biol. 2006;172:803–808. doi: 10.1083/jcb.200512082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Kedersha N, Cho MR, Li W, Yacono PW, Chen S, Gilks N, Golan DE, Anderson P. Dynamic shuttling of TIA-1 accompanies the recruitment of mRNA to mammalian stress granules. J Cell Biol. 2000;151:1257–1268. doi: 10.1083/jcb.151.6.1257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Kedersha N, Stoecklin G, Ayodele M, Yacono P, Lykke-Andersen J, Fritzler MJ, Scheuner D, Kaufman RJ, Golan DE, Anderson P. Stress granules and processing bodies are dynamically linked sites of mRNP remodeling. J Cell Biol. 2005;169:871–884. doi: 10.1083/jcb.200502088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Lykke-Andersen J, Wagner E. Recruitment and activation of mRNA decay enzymes by two ARE-mediated decay activation domains in the proteins TTP and BRF-1. Genes Dev. 2005;19:351–361. doi: 10.1101/gad.1282305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Fenger-Gron M, Fillman C, Norrild B, Lykke-Andersen J. Multiple processing body factors and the ARE binding protein TTP activate mRNA decapping. Mol Cell. 2005;20:905–915. doi: 10.1016/j.molcel.2005.10.031. [DOI] [PubMed] [Google Scholar]
  • 47.Hau HH, Walsh RJ, Ogilvie RL, Williams DA, Reilly CS, Bohjanen PR. Tristetraprolin recruits functional mRNA decay complexes to ARE sequences. J Cell Biochem. 2007;100:1477–1492. doi: 10.1002/jcb.21130. [DOI] [PubMed] [Google Scholar]
  • 48.Barreau C, Paillard L, Osborne HB. AU-rich elements and associated factors: are there unifying principles? Nucleic Acids Res. 2005;33:7138–7150. doi: 10.1093/nar/gki1012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lai WS, Stumpo DJ, Blackshear PJ. Rapid insulin-stimulated accumulation of an mRNA encoding a proline-rich protein. J Biol Chem. 1990;265:16556–16563. [PubMed] [Google Scholar]
  • 50.Stumpo DJ, Byrd NA, Phillips RS, Ghosh S, Maronpot RR, Castranio T, Meyers EN, Mishina Y, Blackshear PJ. Chorioallantoic fusion defects and embryonic lethality resulting from disruption of Zfp36L1, a gene encoding a CCCH tandem zinc finger protein of the Tristetraprolin family. Mol Cell Biol. 2004;24:6445–6455. doi: 10.1128/MCB.24.14.6445-6455.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Stumpo DJ, Broxmeyer HE, Ward T, Cooper S, Hangoc G, Chung YJ, Shelley WC, Richfield EK, Ray MK, Yoder MC, Aplan PD, Blackshear PJ. Targeted disruption of Zfp36l2, encoding a CCCH tandem zinc finger RNA-binding protein, results in defective hematopoiesis. Blood. 2009;114:2401–2410. doi: 10.1182/blood-2009-04-214619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Blackshear PJ, Phillips RS, Ghosh S, Ramos SB, Richfield EK, Lai WS. Zfp36l3, a rodent X chromosome gene encoding a placenta-specific member of the Tristetraprolin family of CCCH tandem zinc finger proteins. Biol Reprod. 2005;73:297–307. doi: 10.1095/biolreprod.105.040527. [DOI] [PubMed] [Google Scholar]
  • 53.Frederick ED, Ramos SB, Blackshear PJ. A unique C-terminal repeat domain maintains the cytosolic localization of the placenta-specific tristetraprolin family member ZFP36L3. J Biol Chem. 2008;283:14792–14800. doi: 10.1074/jbc.M801234200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.DuBois RN, McLane MW, Ryder K, Lau LF, Nathans D. A growth factor-inducible nuclear protein with a novel cysteine/histidine repetitive sequence. J Biol Chem. 1990;265:19185–19191. [PubMed] [Google Scholar]
  • 55.Ma Q, Herschman HR. A corrected sequence for the predicted protein from the mitogen-inducible TIS11 primary response gene. Oncogene. 1991;6:1277–1278. [PubMed] [Google Scholar]
  • 56.Varnum BC, Lim RW, Sukhatme VP, Herschman HR. Nucleotide sequence of a cDNA encoding TIS11, a message induced in Swiss 3T3 cells by the tumor promoter tetradecanoyl phorbol acetate. Oncogene. 1989;4:119–120. [PubMed] [Google Scholar]
  • 57.Taylor GA, Thompson MJ, Lai WS, Blackshear PJ. Mitogens stimulate the rapid nuclear to cytosolic translocation of tristetraprolin, a potential zinc-finger transcription factor. Mol Endocrinol. 1996;10:140–146. doi: 10.1210/mend.10.2.8825554. [DOI] [PubMed] [Google Scholar]
  • 58.Taylor GA, Thompson MJ, Lai WS, Blackshear PJ. Phosphorylation of tristetraprolin, a potential zinc finger transcription factor, by mitogen stimulation in intact cells and by mitogen-activated protein kinase in vitro. J Biol Chem. 1995;270:13341–13347. doi: 10.1074/jbc.270.22.13341. [DOI] [PubMed] [Google Scholar]
  • 59.Carballo E, Lai WS, Blackshear PJ. Feedback inhibition of macrophage tumor necrosis factor-alpha production by tristetraprolin. Science. 1998;281:1001–1005. doi: 10.1126/science.281.5379.1001. [DOI] [PubMed] [Google Scholar]
  • 60.Cao H, Tuttle JS, Blackshear PJ. Immunological characterization of tristetraprolin as a low abundance, inducible, stable cytosolic protein. J Biol Chem. 2004;279:21489–21499. doi: 10.1074/jbc.M400900200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Blackshear PJ, Lai WS, Kennington EA, Brewer G, Wilson GM, Guan X, Zhou P. Characteristics of the interaction of a synthetic human tristetraprolin tandem zinc finger peptide with AU-rich element-containing RNA substrates. J Biol Chem. 2003;278:19947–19955. doi: 10.1074/jbc.M301290200. [DOI] [PubMed] [Google Scholar]
  • 62.Lai WS, Carballo E, Strum JR, Kennington EA, Phillips RS, Blackshear PJ. Evidence that tristetraprolin binds to AU-rich elements and promotes the deadenylation and destabilization of tumor necrosis factor alpha mRNA. Mol Cell Biol. 1999;19:4311–4323. doi: 10.1128/mcb.19.6.4311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Lai WS, Carballo E, Thorn JM, Kennington EA, Blackshear PJ. Interactions of CCCH zinc finger proteins with mRNA. Binding of tristetraprolin-related zinc finger proteins to Au-rich elements and destabilization of mRNA. J Biol Chem. 2000;275:17827–17837. doi: 10.1074/jbc.M001696200. [DOI] [PubMed] [Google Scholar]
  • 64.Lai WS, Blackshear PJ. Interactions of CCCH zinc finger proteins with mRNA: tristetraprolin-mediated AU-rich element-dependent mRNA degradation can occur in the absence of a poly(A) tail. J Biol Chem. 2001;276:23144–23154. doi: 10.1074/jbc.M100680200. [DOI] [PubMed] [Google Scholar]
  • 65.Lai WS, Kennington EA, Blackshear PJ. Tristetraprolin and its family members can promote the cell-free deadenylation of AU-rich element-containing mRNAs by poly(A) ribonuclease. Mol Cell Biol. 2003;23:3798–3812. doi: 10.1128/MCB.23.11.3798-3812.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Carrick DM, Lai WS, Blackshear PJ. The tandem CCCH zinc finger protein tristetraprolin and its relevance to cytokine mRNA turnover and arthritis. Arthritis Res Ther. 2004;6:248–264. doi: 10.1186/ar1441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Lai WS, Carrick DM, Blackshear PJ. Influence of nonameric AU-rich tristetraprolin-binding sites on mRNA deadenylation and turnover. J Biol Chem. 2005;280:34365–34377. doi: 10.1074/jbc.M506757200. [DOI] [PubMed] [Google Scholar]
  • 68.Franks TM, Lykke-Andersen J. TTP and BRF proteins nucleate processing body formation to silence mRNAs with AU-rich elements. Genes Dev. 2007;21:719–735. doi: 10.1101/gad.1494707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Jing Q, Huang S, Guth S, Zarubin T, Motoyama A, Chen J, Di Padova F, Lin SC, Gram H, Han J. Involvement of microRNA in AU-rich element-mediated mRNA instability. Cell. 2005;120:623–634. doi: 10.1016/j.cell.2004.12.038. [DOI] [PubMed] [Google Scholar]
  • 70.Liu J, Valencia-Sanchez MA, Hannon GJ, Parker R. MicroRNA-dependent localization of targeted mRNAs to mammalian P-bodies. Nat Cell Biol. 2005;7:719–723. doi: 10.1038/ncb1274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Taylor GA, Carballo E, Lee DM, Lai WS, Thompson MJ, Patel DD, Schenkman DI, Gilkeson GS, Broxmeyer HE, Haynes BF, Blackshear PJ. A pathogenetic role for TNF alpha in the syndrome of cachexia, arthritis, and autoimmunity resulting from tristetraprolin (TTP) deficiency. Immunity. 1996;4:445–454. doi: 10.1016/s1074-7613(00)80411-2. [DOI] [PubMed] [Google Scholar]
  • 72.Cheng J, Turksen K, Yu QC, Schreiber H, Teng M, Fuchs E. Cachexia and graft-vs.-host-disease-type skin changes in keratin promoter-driven TNF alpha transgenic mice. Genes Dev. 1992;6:1444–1456. doi: 10.1101/gad.6.8.1444. [DOI] [PubMed] [Google Scholar]
  • 73.Ulich TR, Shin SS, del Castillo J. Haematologic effects of TNF. Res Immunol. 1993;144:347–354. doi: 10.1016/s0923-2494(93)80079-e. [DOI] [PubMed] [Google Scholar]
  • 74.Sheehan KC, Ruddle NH, Schreiber RD. Generation and characterization of hamster monoclonal antibodies that neutralize murine tumor necrosis factors. J Immunol. 1989;142:3884–3893. [PubMed] [Google Scholar]
  • 75.Carballo E, Blackshear PJ. Roles of tumor necrosis factor-alpha receptor subtypes in the pathogenesis of the tristetraprolin-deficiency syndrome. Blood. 2001;98:2389–2395. doi: 10.1182/blood.v98.8.2389. [DOI] [PubMed] [Google Scholar]
  • 76.Carballo E, Lai WS, Blackshear PJ. Evidence that tristetraprolin is a physiological regulator of granulocyte-macrophage colony-stimulating factor messenger RNA deadenylation and stability. Blood. 2000;95:1891–1899. [PubMed] [Google Scholar]
  • 77.Stoecklin G, Ming XF, Looser R, Moroni C. Somatic mRNA turnover mutants implicate tristetraprolin in the interleukin-3 mRNA degradation pathway. Mol Cell Biol. 2000;20:3753–3763. doi: 10.1128/mcb.20.11.3753-3763.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Stoecklin G, Stoeckle P, Lu M, Muehlemann O, Moroni C. Cellular mutants define a common mRNA degradation pathway targeting cytokine AU-rich elements. Rna. 2001;7:1578–1588. [PMC free article] [PubMed] [Google Scholar]
  • 79.Stoecklin G, Gross B, Ming XF, Moroni C. A novel mechanism of tumor suppression by destabilizing AU-rich growth factor mRNA. Oncogene. 2003;22:3554–3561. doi: 10.1038/sj.onc.1206418. [DOI] [PubMed] [Google Scholar]
  • 80.Yu H, Stasinopoulos S, Leedman P, Medcalf RL. Inherent instability of plasminogen activator inhibitor type 2 mRNA is regulated by tristetraprolin. J Biol Chem. 2003;278:13912–13918. doi: 10.1074/jbc.M213027200. [DOI] [PubMed] [Google Scholar]
  • 81.Sawaoka H, Dixon DA, Oates JA, Boutaud O. Tristetraprolin binds to the 3′-untranslated region of cyclooxygenase-2 mRNA. A polyadenylation variant in a cancer cell line lacks the binding site. J Biol Chem. 2003;278:13928–13935. doi: 10.1074/jbc.M300016200. [DOI] [PubMed] [Google Scholar]
  • 82.Briata P, Ilengo C, Corte G, Moroni C, Rosenfeld MG, Chen CY, Gherzi R. The Wnt/beta-catenin-->Pitx2 pathway controls the turnover of Pitx2 and other unstable mRNAs. Mol Cell. 2003;12:1201–1211. doi: 10.1016/s1097-2765(03)00407-6. [DOI] [PubMed] [Google Scholar]
  • 83.Suswam E, Li Y, Zhang X, Gillespie GY, Li X, Shacka JJ, Lu L, Zheng L, King PH. Tristetraprolin down-regulates interleukin-8 and vascular endothelial growth factor in malignant glioma cells. Cancer Res. 2008;68:674–682. doi: 10.1158/0008-5472.CAN-07-2751. [DOI] [PubMed] [Google Scholar]
  • 84.Sze KL, Lui WY, Lee WM. Post-transcriptional regulation of CLMP mRNA is controlled by tristetraprolin in response to TNFalpha via c-Jun N-terminal kinase signalling. Biochem J. 2008;410:575–583. doi: 10.1042/BJ20070901. [DOI] [PubMed] [Google Scholar]
  • 85.Emmons J, Townley-Tilson WH, Deleault KM, Skinner SJ, Gross RH, Whitfield ML, Brooks SA. Identification of TTP mRNA targets in human dendritic cells reveals TTP as a critical regulator of dendritic cell maturation. Rna. 2008;14:888–902. doi: 10.1261/rna.748408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Lin NY, Lin CT, Chen YL, Chang CJ. Regulation of tristetraprolin during differentiation of 3T3-L1 preadipocytes. Febs J. 2007;274:867–878. doi: 10.1111/j.1742-4658.2007.05632.x. [DOI] [PubMed] [Google Scholar]
  • 87.Balakathiresan NS, Bhattacharyya S, Gutti U, Long RP, Jozwik C, Huang W, Srivastava M, Pollard HB, Biswas R. Tristetraprolin regulates IL-8 mRNA stability in cystic fibrosis lung epithelial cells. Am J Physiol Lung Cell Mol Physiol. 2009;296:L1012–1018. doi: 10.1152/ajplung.90601.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Essafi-Benkhadir K, Onesto C, Stebe E, Moroni C, Pages G. Tristetraprolin inhibits Ras-dependent tumor vascularization by inducing vascular endothelial growth factor mRNA degradation. Mol Biol Cell. 2007;18:4648–4658. doi: 10.1091/mbc.E07-06-0570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Frasca D, Landin AM, Alvarez JP, Blackshear PJ, Riley RL, Blomberg BB. Tristetraprolin, a negative regulator of mRNA stability, is increased in old B cells and is involved in the degradation of E47 mRNA. J Immunol. 2007;179:918–927. doi: 10.4049/jimmunol.179.2.918. [DOI] [PubMed] [Google Scholar]
  • 90.Patino WD, Kang JG, Matoba S, Mian OY, Gochuico BR, Hwang PM. Atherosclerotic plaque macrophage transcriptional regulators are expressed in blood and modulated by tristetraprolin. Circ Res. 2006;98:1282–1289. doi: 10.1161/01.RES.0000222284.48288.28. [DOI] [PubMed] [Google Scholar]
  • 91.Jalonen U, Nieminen R, Vuolteenaho K, Kankaanranta H, Moilanen E. Down-regulation of tristetraprolin expression results in enhanced IL-12 and MIP-2 production and reduced MIP-3alpha synthesis in activated macrophages. Mediators Inflamm. 2006;2006:40691. doi: 10.1155/MI/2006/40691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Marderosian M, Sharma A, Funk AP, Vartanian R, Masri J, Jo OD, Gera JF. Tristetraprolin regulates Cyclin D1 and c-Myc mRNA stability in response to rapamycin in an Akt-dependent manner via p38 MAPK signaling. Oncogene. 2006;25:6277–6290. doi: 10.1038/sj.onc.1209645. [DOI] [PubMed] [Google Scholar]
  • 93.Gringhuis SI, Garcia-Vallejo JJ, van Het Hof B, van Dijk W. Convergent actions of I kappa B kinase beta and protein kinase C delta modulate mRNA stability through phosphorylation of 14-3-3 beta complexed with tristetraprolin. Mol Cell Biol. 2005;25:6454–6463. doi: 10.1128/MCB.25.15.6454-6463.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Fechir M, Linker K, Pautz A, Hubrich T, Forstermann U, Rodriguez-Pascual F, Kleinert H. Tristetraprolin regulates the expression of the human inducible nitric-oxide synthase gene. Mol Pharmacol. 2005;67:2148–2161. doi: 10.1124/mol.104.008763. [DOI] [PubMed] [Google Scholar]
  • 95.Linker K, Pautz A, Fechir M, Hubrich T, Greeve J, Kleinert H. Involvement of KSRP in the post-transcriptional regulation of human iNOS expression-complex interplay of KSRP with TTP and HuR. Nucleic Acids Res. 2005;33:4813–4827. doi: 10.1093/nar/gki797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Ogilvie RL, Abelson M, Hau HH, Vlasova I, Blackshear PJ, Bohjanen PR. Tristetraprolin down-regulates IL-2 gene expression through AU-rich element-mediated mRNA decay. J Immunol. 2005;174:953–961. doi: 10.4049/jimmunol.174.2.953. [DOI] [PubMed] [Google Scholar]
  • 97.Ogilvie RL, Sternjohn JR, Rattenbacher B, Vlasova IA, Williams DA, Hau HH, Blackshear PJ, Bohjanen PR. Tristetraprolin mediates interferon-gamma mRNA decay. J Biol Chem. 2009;284:11216–11223. doi: 10.1074/jbc.M901229200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Romagnani P, Lasagni L, Annunziato F, Serio M, Romagnani S. CXC chemokines: the regulatory link between inflammation and angiogenesis. Trends Immunol. 2004;25:201–209. doi: 10.1016/j.it.2004.02.006. [DOI] [PubMed] [Google Scholar]
  • 99.Datta S, Biswas R, Novotny M, Pavicic PG, Jr, Herjan T, Mandal P, Hamilton TA. Tristetraprolin regulates CXCL1 (KC) mRNA stability. J Immunol. 2008;180:2545–2552. doi: 10.4049/jimmunol.180.4.2545. [DOI] [PubMed] [Google Scholar]
  • 100.Novotny M, Datta S, Biswas R, Hamilton T. Functionally independent AU-rich sequence motifs regulate KC (CXCL1) mRNA. J Biol Chem. 2005;280:30166–30174. doi: 10.1074/jbc.M502280200. [DOI] [PubMed] [Google Scholar]
  • 101.Lai WS, Parker JS, Grissom SF, Stumpo DJ, Blackshear PJ. Novel mRNA targets for tristetraprolin (TTP) identified by global analysis of stabilized transcripts in TTP-deficient fibroblasts. Mol Cell Biol. 2006;26:9196–9208. doi: 10.1128/MCB.00945-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Horner TJ, Lai WS, Stumpo DJ, Blackshear PJ. Stimulation of polo-like kinase 3 mRNA decay by tristetraprolin. Mol Cell Biol. 2009;29:1999–2010. doi: 10.1128/MCB.00982-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Stoecklin G, Tenenbaum SA, Mayo T, Chittur SV, George AD, Baroni TE, Blackshear PJ, Anderson P. Genome-wide analysis identifies interleukin-10 mRNA as target of tristetraprolin. J Biol Chem. 2008;283:11689–11699. doi: 10.1074/jbc.M709657200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Sommer SL, Berndt TJ, Frank E, Patel JB, Redfield MM, Dong X, Griffin MD, Grande JP, van Deursen JM, Sieck GC, Romero JC, Kumar R. Elevated blood pressure and cardiac hypertrophy after ablation of the gly96/IEX-1 gene. J Appl Physiol. 2006;100:707–716. doi: 10.1152/japplphysiol.00306.2005. [DOI] [PubMed] [Google Scholar]
  • 105.Xie S, Wu H, Wang Q, Cogswell JP, Husain I, Conn C, Stambrook P, Jhanwar-Uniyal M, Dai W. Plk3 functionally links DNA damage to cell cycle arrest and apoptosis at least in part via the p53 pathway. J Biol Chem. 2001;276:43305–43312. doi: 10.1074/jbc.M106050200. [DOI] [PubMed] [Google Scholar]
  • 106.Winkles JA, Alberts GF. Differential regulation of polo-like kinase 1, 2, 3, and 4 gene expression in mammalian cells and tissues. Oncogene. 2005;24:260–266. doi: 10.1038/sj.onc.1208219. [DOI] [PubMed] [Google Scholar]
  • 107.Zimmerman WC, Erikson RL. Polo-like kinase 3 is required for entry into S phase. Proc Natl Acad Sci U S A. 2007;104:1847–1852. doi: 10.1073/pnas.0610856104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Bell SE, Sanchez MJ, Spasic-Boskovic O, Santalucia T, Gambardella L, Burton GJ, Murphy JJ, Norton JD, Clark AR, Turner M. The RNA binding protein Zfp36l1 is required for normal vascularisation and post-transcriptionally regulates VEGF expression. Dev Dyn. 2006;235:3144–3155. doi: 10.1002/dvdy.20949. [DOI] [PubMed] [Google Scholar]
  • 109.Ciais D, Cherradi N, Bailly S, Grenier E, Berra E, Pouyssegur J, Lamarre J, Feige JJ. Destabilization of vascular endothelial growth factor mRNA by the zinc-finger protein TIS11b. Oncogene. 2004;23:8673–8680. doi: 10.1038/sj.onc.1207939. [DOI] [PubMed] [Google Scholar]
  • 110.Stoecklin G, Colombi M, Raineri I, Leuenberger S, Mallaun M, Schmidlin M, Gross B, Lu M, Kitamura T, Moroni C. Functional cloning of BRF1, a regulator of ARE-dependent mRNA turnover. Embo J. 2002;21:4709–4718. doi: 10.1093/emboj/cdf444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Duan H, Cherradi N, Feige JJ, Jefcoate C. cAMP-dependent posttranscriptional regulation of steroidogenic acute regulatory (STAR) protein by the zinc finger protein ZFP36L1/TIS11b. Mol Endocrinol. 2009;23:497–509. doi: 10.1210/me.2008-0296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Ramos SB, Stumpo DJ, Kennington EA, Phillips RS, Bock CB, Ribeiro-Neto F, Blackshear PJ. The CCCH tandem zinc-finger protein Zfp36l2 is crucial for female fertility and early embryonic development. Development. 2004;131:4883–4893. doi: 10.1242/dev.01336. [DOI] [PubMed] [Google Scholar]
  • 113.Johnson BA, Stehn JR, Yaffe MB, Blackwell TK. Cytoplasmic localization of tristetraprolin involves 14-3-3-dependent and -independent mechanisms. J Biol Chem. 2002;277:18029–18036. doi: 10.1074/jbc.M110465200. [DOI] [PubMed] [Google Scholar]
  • 114.Brewer G, Ross J. Regulation of c-myc mRNA stability in vitro by a labile destabilizer with an essential nucleic acid component. Mol Cell Biol. 1989;9:1996–2006. doi: 10.1128/mcb.9.5.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Brewer G. An A + U-rich element RNA-binding factor regulates c-myc mRNA stability in vitro. Mol Cell Biol. 1991;11:2460–2466. doi: 10.1128/mcb.11.5.2460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Zhang W, Wagner BJ, Ehrenman K, Schaefer AW, DeMaria CT, Crater D, DeHaven K, Long L, Brewer G. Purification, characterization, and cDNA cloning of an AU-rich element RNA-binding protein, AUF1. Mol Cell Biol. 1993;13:7652–7665. doi: 10.1128/mcb.13.12.7652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Loflin P, Chen CY, Shyu AB. Unraveling a cytoplasmic role for hnRNP D in the in vivo mRNA destabilization directed by the AU-rich element. Genes Dev. 1999;13:1884–1897. doi: 10.1101/gad.13.14.1884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Raineri I, Wegmueller D, Gross B, Certa U, Moroni C. Roles of AUF1 isoforms, HuR and BRF1 in ARE-dependent mRNA turnover studied by RNA interference. Nucleic Acids Res. 2004;32:1279–1288. doi: 10.1093/nar/gkh282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Lal A, Mazan-Mamczarz K, Kawai T, Yang X, Martindale JL, Gorospe M. Concurrent versus individual binding of HuR and AUF1 to common labile target mRNAs. Embo J. 2004;23:3092–3102. doi: 10.1038/sj.emboj.7600305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Pautz A, Linker K, Altenhofer S, Heil S, Schmidt N, Art J, Knauer S, Stauber R, Sadri N, Pont A, Schneider RJ, Kleinert H. Similar regulation of human inducible nitric-oxide synthase expression by different isoforms of the RNA-binding protein AUF1. J Biol Chem. 2009;284:2755–2766. doi: 10.1074/jbc.M809314200. [DOI] [PubMed] [Google Scholar]
  • 121.Xu N, Chen CY, Shyu AB. Modulation of the fate of cytoplasmic mRNA by AU-rich elements: key sequence features controlling mRNA deadenylation and decay. Mol Cell Biol. 1997;17:4611–4621. doi: 10.1128/mcb.17.8.4611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Sela-Brown A, Silver J, Brewer G, Naveh-Many T. Identification of AUF1 as a parathyroid hormone mRNA 3′-untranslated region-binding protein that determines parathyroid hormone mRNA stability. J Biol Chem. 2000;275:7424–7429. doi: 10.1074/jbc.275.10.7424. [DOI] [PubMed] [Google Scholar]
  • 123.Wagner BJ, DeMaria CT, Sun Y, Wilson GM, Brewer G. Structure and genomic organization of the human AUF1 gene: alternative pre-mRNA splicing generates four protein isoforms. Genomics. 1998;48:195–202. doi: 10.1006/geno.1997.5142. [DOI] [PubMed] [Google Scholar]
  • 124.Buzby JS, Lee SM, Van Winkle P, DeMaria CT, Brewer G, Cairo MS. Increased granulocyte-macrophage colony-stimulating factor mRNA instability in cord versus adult mononuclear cells is translation-dependent and associated with increased levels of A + U-rich element binding factor. Blood. 1996;88:2889–2897. [PubMed] [Google Scholar]
  • 125.Sirenko OI, Lofquist AK, DeMaria CT, Morris JS, Brewer G, Haskill JS. Adhesion-dependent regulation of an A+U-rich element-binding activity associated with AUF1. Mol Cell Biol. 1997;17:3898–3906. doi: 10.1128/mcb.17.7.3898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Sarkar B, Xi Q, He C, Schneider RJ. Selective degradation of AU-rich mRNAs promoted by the p37 AUF1 protein isoform. Mol Cell Biol. 2003;23:6685–6693. doi: 10.1128/MCB.23.18.6685-6693.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Shyu AB, Wilkinson MF. The double lives of shuttling mRNA binding proteins. Cell. 2000;102:135–138. doi: 10.1016/s0092-8674(00)00018-0. [DOI] [PubMed] [Google Scholar]
  • 128.Sarkar B, Lu JY, Schneider RJ. Nuclear import and export functions in the different isoforms of the AUF1/heterogeneous nuclear ribonucleoprotein protein family. J Biol Chem. 2003;278:20700–20707. doi: 10.1074/jbc.M301176200. [DOI] [PubMed] [Google Scholar]
  • 129.Lu JY, Sadri N, Schneider RJ. Endotoxic shock in AUF1 knockout mice mediated by failure to degrade proinflammatory cytokine mRNAs. Genes Dev. 2006;20:3174–3184. doi: 10.1101/gad.1467606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Sadri N, Schneider RJ. Auf1/Hnrnpd-deficient mice develop pruritic inflammatory skin disease. J Invest Dermatol. 2009;129:657–670. doi: 10.1038/jid.2008.298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Ma WJ, Cheng S, Campbell C, Wright A, Furneaux H. Cloning and characterization of HuR, a ubiquitously expressed Elav-like protein. J Biol Chem. 1996;271:8144–8151. doi: 10.1074/jbc.271.14.8144. [DOI] [PubMed] [Google Scholar]
  • 132.Good PJ. A conserved family of elav-like genes in vertebrates. Proc Natl Acad Sci U S A. 1995;92:4557–4561. doi: 10.1073/pnas.92.10.4557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Antic D, Keene JD. Embryonic lethal abnormal visual RNA-binding proteins involved in growth, differentiation, and posttranscriptional gene expression. Am J Hum Genet. 1997;61:273–278. doi: 10.1086/514866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Fan XC, Myer VE, Steitz JA. AU-rich elements target small nuclear RNAs as well as mRNAs for rapid degradation. Genes Dev. 1997;11:2557–2568. doi: 10.1101/gad.11.19.2557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Fan XC, Steitz JA. HNS, a nuclear-cytoplasmic shuttling sequence in HuR. Proc Natl Acad Sci U S A. 1998;95:15293–15298. doi: 10.1073/pnas.95.26.15293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Dean JL, Wait R, Mahtani KR, Sully G, Clark AR, Saklatvala J. The 3′ untranslated region of tumor necrosis factor alpha mRNA is a target of the mRNA-stabilizing factor HuR. Mol Cell Biol. 2001;21:721–730. doi: 10.1128/MCB.21.3.721-730.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Levy NS, Chung S, Furneaux H, Levy AP. Hypoxic stabilization of vascular endothelial growth factor mRNA by the RNA-binding protein HuR. J Biol Chem. 1998;273:6417–6423. doi: 10.1074/jbc.273.11.6417. [DOI] [PubMed] [Google Scholar]
  • 138.Nabors LB, Suswam E, Huang Y, Yang X, Johnson MJ, King PH. Tumor necrosis factor alpha induces angiogenic factor up-regulation in malignant glioma cells: a role for RNA stabilization and HuR. Cancer Res. 2003;63:4181–4187. [PubMed] [Google Scholar]
  • 139.Dixon DA, Tolley ND, King PH, Nabors LB, McIntyre TM, Zimmerman GA, Prescott SM. Altered expression of the mRNA stability factor HuR promotes cyclooxygenase-2 expression in colon cancer cells. J Clin Invest. 2001;108:1657–1665. doi: 10.1172/JCI12973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Peng SS, Chen CY, Xu N, Shyu AB. RNA stabilization by the AU-rich element binding protein, HuR, an ELAV protein. Embo J. 1998;17:3461–3470. doi: 10.1093/emboj/17.12.3461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Fan XC, Steitz JA. Overexpression of HuR, a nuclear-cytoplasmic shuttling protein, increases the in vivo stability of ARE-containing mRNAs. Embo J. 1998;17:3448–3460. doi: 10.1093/emboj/17.12.3448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Yang X, Wang W, Fan J, Lal A, Yang D, Cheng H, Gorospe M. Prostaglandin A2-mediated stabilization of p21 mRNA through an ERK-dependent pathway requiring the RNA-binding protein HuR. J Biol Chem. 2004;279:49298–49306. doi: 10.1074/jbc.M407535200. [DOI] [PubMed] [Google Scholar]
  • 143.Wang W, Caldwell MC, Lin S, Furneaux H, Gorospe M. HuR regulates cyclin A and cyclin B1 mRNA stability during cell proliferation. Embo J. 2000;19:2340–2350. doi: 10.1093/emboj/19.10.2340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Rodriguez-Pascual F, Hausding M, Ihrig-Biedert I, Furneaux H, Levy AP, Forstermann U, Kleinert H. Complex contribution of the 3′-untranslated region to the expressional regulation of the human inducible nitric-oxide synthase gene. Involvement of the RNA-binding protein HuR. J Biol Chem. 2000;275:26040–26049. doi: 10.1074/jbc.M910460199. [DOI] [PubMed] [Google Scholar]
  • 145.Xiao L, Rao JN, Zou T, Liu L, Marasa BS, Chen J, Turner DJ, Zhou H, Gorospe M, Wang JY. Polyamines regulate the stability of activating transcription factor-2 mRNA through RNA-binding protein HuR in intestinal epithelial cells. Mol Biol Cell. 2007;18:4579–4590. doi: 10.1091/mbc.E07-07-0675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Zhang X, Zou T, Rao JN, Liu L, Xiao L, Wang PY, Cui YH, Gorospe M, Wang JY. Stabilization of XIAP mRNA through the RNA binding protein HuR regulated by cellular polyamines. Nucleic Acids Res. 2009 doi: 10.1093/nar/gkp755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Mazan-Mamczarz K, Galban S, Lopez de Silanes I, Martindale JL, Atasoy U, Keene JD, Gorospe M. RNA-binding protein HuR enhances p53 translation in response to ultraviolet light irradiation. Proc Natl Acad Sci U S A. 2003;100:8354–8359. doi: 10.1073/pnas.1432104100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Galban S, Martindale JL, Mazan-Mamczarz K, Lopez de Silanes I, Fan J, Wang W, Decker J, Gorospe M. Influence of the RNA-binding protein HuR in pVHL-regulated p53 expression in renal carcinoma cells. Mol Cell Biol. 2003;23:7083–7095. doi: 10.1128/MCB.23.20.7083-7095.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Galban S, Kuwano Y, Pullmann R, Jr, Martindale JL, Kim HH, Lal A, Abdelmohsen K, Yang X, Dang Y, Liu JO, Lewis SM, Holcik M, Gorospe M. RNA-binding proteins HuR and PTB promote the translation of hypoxia-inducible factor 1alpha. Mol Cell Biol. 2008;28:93–107. doi: 10.1128/MCB.00973-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Katsanou V, Papadaki O, Milatos S, Blackshear PJ, Anderson P, Kollias G, Kontoyiannis DL. HuR as a negative posttranscriptional modulator in inflammation. Mol Cell. 2005;19:777–789. doi: 10.1016/j.molcel.2005.08.007. [DOI] [PubMed] [Google Scholar]
  • 151.Katsanou V, Milatos S, Yiakouvaki A, Sgantzis N, Kotsoni A, Alexiou M, Harokopos V, Aidinis V, Hemberger M, Kontoyiannis DL. The RNA-binding protein Elavl1/HuR is essential for placental branching morphogenesis and embryonic development. Mol Cell Biol. 2009;29:2762–2776. doi: 10.1128/MCB.01393-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Ghosh M, Aguila HL, Michaud J, Ai Y, Wu MT, Hemmes A, Ristimaki A, Guo C, Furneaux H, Hla T. Essential role of the RNA-binding protein HuR in progenitor cell survival in mice. J Clin Invest. 2009 doi: 10.1172/JCI38263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Tian Q, Streuli M, Saito H, Schlossman SF, Anderson P. A polyadenylate binding protein localized to the granules of cytolytic lymphocytes induces DNA fragmentation in target cells. Cell. 1991;67:629–639. doi: 10.1016/0092-8674(91)90536-8. [DOI] [PubMed] [Google Scholar]
  • 154.Kawakami A, Tian Q, Duan X, Streuli M, Schlossman SF, Anderson P. Identification and functional characterization of a TIA-1-related nucleolysin. Proc Natl Acad Sci U S A. 1992;89:8681–8685. doi: 10.1073/pnas.89.18.8681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Beck AR, Medley QG, O’Brien S, Anderson P, Streuli M. Structure, tissue distribution and genomic organization of the murine RRM-type RNA binding proteins TIA-1 and TIAR. Nucleic Acids Res. 1996;24:3829–3835. doi: 10.1093/nar/24.19.3829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Dember LM, Kim ND, Liu KQ, Anderson P. Individual RNA recognition motifs of TIA-1 and TIAR have different RNA binding specificities. J Biol Chem. 1996;271:2783–2788. doi: 10.1074/jbc.271.5.2783. [DOI] [PubMed] [Google Scholar]
  • 157.Gueydan C, Droogmans L, Chalon P, Huez G, Caput D, Kruys V. Identification of TIAR as a protein binding to the translational regulatory AU-rich element of tumor necrosis factor alpha mRNA. J Biol Chem. 1999;274:2322–2326. doi: 10.1074/jbc.274.4.2322. [DOI] [PubMed] [Google Scholar]
  • 158.Piecyk M, Wax S, Beck AR, Kedersha N, Gupta M, Maritim B, Chen S, Gueydan C, Kruys V, Streuli M, Anderson P. TIA-1 is a translational silencer that selectively regulates the expression of TNF-alpha. Embo J. 2000;19:4154–4163. doi: 10.1093/emboj/19.15.4154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Dixon DA, Balch GC, Kedersha N, Anderson P, Zimmerman GA, Beauchamp RD, Prescott SM. Regulation of cyclooxygenase-2 expression by the translational silencer TIA-1. J Exp Med. 2003;198:475–481. doi: 10.1084/jem.20030616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Cok SJ, Acton SJ, Morrison AR. The proximal region of the 3′-untranslated region of cyclooxygenase-2 is recognized by a multimeric protein complex containing HuR, TIA-1, TIAR, and the heterogeneous nuclear ribonucleoprotein U. J Biol Chem. 2003;278:36157–36162. doi: 10.1074/jbc.M302547200. [DOI] [PubMed] [Google Scholar]
  • 161.Phillips K, Kedersha N, Shen L, Blackshear PJ, Anderson P. Arthritis suppressor genes TIA-1 and TTP dampen the expression of tumor necrosis factor alpha, cyclooxygenase 2, and inflammatory arthritis. Proc Natl Acad Sci U S A. 2004;101:2011–2016. doi: 10.1073/pnas.0400148101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Yu Q, Cok SJ, Zeng C, Morrison AR. Translational repression of human matrix metalloproteinases-13 by an alternatively spliced form of T-cell-restricted intracellular antigen-related protein (TIAR) J Biol Chem. 2003;278:1579–1584. doi: 10.1074/jbc.M203526200. [DOI] [PubMed] [Google Scholar]
  • 163.Saito K, Chen S, Piecyk M, Anderson P. TIA-1 regulates the production of tumor necrosis factor alpha in macrophages, but not in lymphocytes. Arthritis Rheum. 2001;44:2879–2887. doi: 10.1002/1529-0131(200112)44:12<2879::aid-art476>3.0.co;2-4. [DOI] [PubMed] [Google Scholar]
  • 164.Kedersha N, Anderson P. Stress granules: sites of mRNA triage that regulate mRNA stability and translatability. Biochem Soc Trans. 2002;30:963–969. doi: 10.1042/bst0300963. [DOI] [PubMed] [Google Scholar]
  • 165.Anderson P, Kedersha N. Stressful initiations. J Cell Sci. 2002;115:3227–3234. doi: 10.1242/jcs.115.16.3227. [DOI] [PubMed] [Google Scholar]
  • 166.Anant S, MacGinnitie AJ, Davidson NO. apobec-1, the catalytic subunit of the mammalian apolipoprotein B mRNA editing enzyme, is a novel RNA-binding protein. J Biol Chem. 1995;270:14762–14767. [PubMed] [Google Scholar]
  • 167.MacGinnitie AJ, Anant S, Davidson NO. Mutagenesis of apobec-1, the catalytic subunit of the mammalian apolipoprotein B mRNA editing enzyme, reveals distinct domains that mediate cytosine nucleoside deaminase, RNA binding, and RNA editing activity. J Biol Chem. 1995;270:14768–14775. [PubMed] [Google Scholar]
  • 168.Anant S, Davidson NO. An AU-rich sequence element (UUUN[A/U]U) downstream of the edited C in apolipoprotein B mRNA is a high-affinity binding site for Apobec-1: binding of Apobec-1 to this motif in the 3′ untranslated region of c-myc increases mRNA stability. Mol Cell Biol. 2000;20:1982–1992. doi: 10.1128/mcb.20.6.1982-1992.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Anant S, Murmu N, Houchen CW, Mukhopadhyay D, Riehl TE, Young SG, Morrison AR, Stenson WF, Davidson NO. Apobec-1 protects intestine from radiation injury through posttranscriptional regulation of cyclooxygenase-2 expression. Gastroenterology. 2004;127:1139–1149. doi: 10.1053/j.gastro.2004.06.022. [DOI] [PubMed] [Google Scholar]
  • 170.Dean JL, Sully G, Wait R, Rawlinson L, Clark AR, Saklatvala J. Identification of a novel AU-rich-element-binding protein which is related to AUF1. Biochem J. 2002;366:709–719. doi: 10.1042/BJ20020402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Subramaniam D, Natarajan G, Ramalingam S, Ramachandran I, May R, Queimado L, Houchen CW, Anant S. Translation inhibition during cell cycle arrest and apoptosis: Mcl-1 is a novel target for RNA binding protein CUGBP2. Am J Physiol Gastrointest Liver Physiol. 2008;294:G1025–1032. doi: 10.1152/ajpgi.00602.2007. [DOI] [PubMed] [Google Scholar]
  • 172.Mukhopadhyay D, Houchen CW, Kennedy S, Dieckgraefe BK, Anant S. Coupled mRNA stabilization and translational silencing of cyclooxygenase-2 by a novel RNA binding protein, CUGBP2. Mol Cell. 2003;11:113–126. doi: 10.1016/s1097-2765(03)00012-1. [DOI] [PubMed] [Google Scholar]
  • 173.Garnon J, Lachance C, Di Marco S, Hel Z, Marion D, Ruiz MC, Newkirk MM, Khandjian EW, Radzioch D. Fragile X-related protein FXR1P regulates proinflammatory cytokine tumor necrosis factor expression at the post-transcriptional level. J Biol Chem. 2005;280:5750–5763. doi: 10.1074/jbc.M401988200. [DOI] [PubMed] [Google Scholar]
  • 174.Gherzi R, Lee KY, Briata P, Wegmuller D, Moroni C, Karin M, Chen CY. A KH domain RNA binding protein, KSRP, promotes ARE-directed mRNA turnover by recruiting the degradation machinery. Mol Cell. 2004;14:571–583. doi: 10.1016/j.molcel.2004.05.002. [DOI] [PubMed] [Google Scholar]
  • 175.Gherzi R, Trabucchi M, Ponassi M, Ruggiero T, Corte G, Moroni C, Chen CY, Khabar KS, Andersen JS, Briata P. The RNA-binding protein KSRP promotes decay of beta-catenin mRNA and is inactivated by PI3K-AKT signaling. PLoS Biol. 2006;5:e5. doi: 10.1371/journal.pbio.0050005. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 176.Winzen R, Thakur BK, Dittrich-Breiholz O, Shah M, Redich N, Dhamija S, Kracht M, Holtmann H. Functional analysis of KSRP interaction with the AU-rich element of interleukin-8 and identification of inflammatory mRNA targets. Mol Cell Biol. 2007;27:8388–8400. doi: 10.1128/MCB.01493-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Nechama M, Ben-Dov IZ, Briata P, Gherzi R, Naveh-Many T. The mRNA decay promoting factor K-homology splicing regulator protein post-transcriptionally determines parathyroid hormone mRNA levels. Faseb J. 2008;22:3458–3468. doi: 10.1096/fj.08-107250. [DOI] [PubMed] [Google Scholar]
  • 178.Nechama M, Peng Y, Bell O, Briata P, Gherzi R, Schoenberg DR, Naveh-Many T. KSRP-PMR1-exosome association determines parathyroid hormone mRNA levels and stability in transfected cells. BMC Cell Biol. 2009;10:70. doi: 10.1186/1471-2121-10-70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Shim J, Lim H, JRY, Karin M. Nuclear export of NF90 is required for interleukin-2 mRNA stabilization. Mol Cell. 2002;10:1331–1344. doi: 10.1016/s1097-2765(02)00730-x. [DOI] [PubMed] [Google Scholar]
  • 180.Shi L, Zhao G, Qiu D, Godfrey WR, Vogel H, Rando TA, Hu H, Kao PN. NF90 regulates cell cycle exit and terminal myogenic differentiation by direct binding to the 3′-untranslated region of MyoD and p21WAF1/CIP1 mRNAs. J Biol Chem. 2005;280:18981–18989. doi: 10.1074/jbc.M411034200. [DOI] [PubMed] [Google Scholar]
  • 181.Shi L, Godfrey WR, Lin J, Zhao G, Kao PN. NF90 regulates inducible IL-2 gene expression in T cells. J Exp Med. 2007;204:971–977. doi: 10.1084/jem.20052078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Vumbaca F, Phoenix KN, Rodriguez-Pinto D, Han DK, Claffey KP. Double-stranded RNA-binding protein regulates vascular endothelial growth factor mRNA stability, translation, and breast cancer angiogenesis. Mol Cell Biol. 2008;28:772–783. doi: 10.1128/MCB.02078-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Kuwano Y, Kim HH, Abdelmohsen K, Pullmann R, Jr, Martindale JL, Yang X, Gorospe M. MKP-1 mRNA stabilization and translational control by RNA-binding proteins HuR and NF90. Mol Cell Biol. 2008;28:4562–4575. doi: 10.1128/MCB.00165-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Kuwano Y, Pullmann R, Jr, Marasa BS, Abdelmohsen K, Lee EK, Yang X, Martindale JL, Zhan M, Gorospe M. NF90 selectively represses the translation of target mRNAs bearing an AU-rich signature motif. Nucleic Acids Res. 2009 doi: 10.1093/nar/gkp861. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Tran H, Schilling M, Wirbelauer C, Hess D, Nagamine Y. Facilitation of mRNA deadenylation and decay by the exosome-bound, DExH protein RHAU. Mol Cell. 2004;13:101–111. doi: 10.1016/s1097-2765(03)00481-7. [DOI] [PubMed] [Google Scholar]
  • 186.Chalupnikova K, Lattmann S, Selak N, Iwamoto F, Fujiki Y, Nagamine Y. Recruitment of the RNA helicase RHAU to stress granules via a unique RNA-binding domain. J Biol Chem. 2008;283:35186–35198. doi: 10.1074/jbc.M804857200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Capowski EE, Esnault S, Bhattacharya S, Malter JS. Y box-binding factor promotes eosinophil survival by stabilizing granulocyte-macrophage colony-stimulating factor mRNA. J Immunol. 2001;167:5970–5976. doi: 10.4049/jimmunol.167.10.5970. [DOI] [PubMed] [Google Scholar]
  • 188.Coles LS, Bartley MA, Bert A, Hunter J, Polyak S, Diamond P, Vadas MA, Goodall GJ. A multi-protein complex containing cold shock domain (Y-box) and polypyrimidine tract binding proteins forms on the vascular endothelial growth factor mRNA. Potential role in mRNA stabilization. Eur J Biochem. 2004;271:648–660. doi: 10.1111/j.1432-1033.2003.03968.x. [DOI] [PubMed] [Google Scholar]
  • 189.Kandasamy K, Joseph K, Subramaniam K, Raymond JR, Tholanikunnel BG. Translational control of beta2-adrenergic receptor mRNA by T-cell-restricted intracellular antigen-related protein. J Biol Chem. 2005;280:1931–1943. doi: 10.1074/jbc.M405937200. [DOI] [PubMed] [Google Scholar]
  • 190.Mazan-Mamczarz K, Lal A, Martindale JL, Kawai T, Gorospe M. Translational repression by RNA-binding protein TIAR. Mol Cell Biol. 2006;26:2716–2727. doi: 10.1128/MCB.26.7.2716-2727.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Tudor C, Marchese FP, Hitti E, Aubareda A, Rawlinson L, Gaestel M, Blackshear PJ, Clark AR, Saklatvala J, Dean JL. The p38 MAPK pathway inhibits tristetraprolin-directed decay of interleukin-10 and pro-inflammatory mediator mRNAs in murine macrophages. FEBS Lett. 2009;583:1933–1938. doi: 10.1016/j.febslet.2009.04.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Carballo E, Gilkeson GS, Blackshear PJ. Bone marrow transplantation reproduces the tristetraprolin-deficiency syndrome in recombination activating gene-2 (−/−) mice. Evidence that monocyte/macrophage progenitors may be responsible for TNFalpha overproduction. J Clin Invest. 1997;100:986–995. doi: 10.1172/JCI119649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Lai WS, Kennington EA, Blackshear PJ. Interactions of CCCH zinc finger proteins with mRNA: non-binding tristetraprolin mutants exert an inhibitory effect on degradation of AU-rich element-containing mRNAs. J Biol Chem. 2002;277:9606–9613. doi: 10.1074/jbc.M110395200. [DOI] [PubMed] [Google Scholar]
  • 194.Sauer I, Schaljo B, Vogl C, Gattermeier I, Kolbe T, Muller M, Blackshear PJ, Kovarik P. Interferons limit inflammatory responses by induction of tristetraprolin. Blood. 2006;107:4790–4797. doi: 10.1182/blood-2005-07-3058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Lee SK, Kim SB, Kim JS, Moon CH, Han MS, Lee BJ, Chung DK, Min YJ, Park JH, Choi DH, Cho HR, Park SK, Park JW. Butyrate response factor 1 enhances cisplatin sensitivity in human head and neck squamous cell carcinoma cell lines. Int J Cancer. 2005;117:32–40. doi: 10.1002/ijc.21133. [DOI] [PubMed] [Google Scholar]

RESOURCES