The AP-1 transcription factor is an inducible dimeric protein complex that regulates gene expression by binding to specific DNA sequences, such as the TPA-responsive element (TRE; 5′-TGAG/CTCA-3′) or cAMP-responsive element (CRE; 5′-TGACGTCA-3′), in response to extracellular stimuli.[1][2] It consists of homo- or heterodimers formed by basic leucine zipper (bZIP) domain-containing proteins from the Jun (c-Jun, JunB, JunD), Fos (c-Fos, FosB, Fra-1, Fra-2), ATF (ATF2, ATF3, BATF), and Maf (c-Maf, MafA, MafB, MafG/F/K) families, enabling diverse functional specificities.[1][2] First identified in the 1980s as a DNA-binding activity in mammalian cells and linked to proto-oncogenes like c-jun and c-fos discovered in retroviruses, AP-1 was formally characterized in the late 1980s through studies on phorbol ester-induced gene activation.[1][2]AP-1's activity is primarily regulated by mitogen-activated protein kinase (MAPK) pathways, including ERK, JNK, and p38, which phosphorylate its subunits to enhance DNA binding, dimerization, and transactivation potential, often in concert with NF-κB signaling.[1][2] Additional control occurs through post-translational modifications, protein interactions, and transcriptional regulation of subunit expression, allowing context-specific responses to growth factors, cytokines, and stress signals.[1] In physiological contexts, AP-1 governs essential cellular processes such as proliferation, differentiation, survival, apoptosis, and migration, with particular importance in immune cell activation (e.g., T-cell cytokine production like IL-2 and IL-6) and epidermal keratinocytehomeostasis.[1][2]Dysregulation of AP-1 contributes to pathological conditions, exhibiting dual roles in cancer where subunits like c-Jun promote oncogenesis via proliferation and invasion, while JunB and JunD often act as tumor suppressors.[1] It also drives inflammatory responses and immune evasion mechanisms, such as upregulation of checkpoints like PD-1 and PD-L1 in tumors.[2] These multifaceted properties position AP-1 as a key therapeutic target, with inhibitors explored for anticancer and anti-inflammatory applications.[1]
Discovery and History
Initial Identification
The activator protein 1 (AP-1) transcription factor was first identified in the mid-1980s as a DNA-binding activity in nuclear extracts from phorbol ester (12-O-tetradecanoylphorbol-13-acetate, or TPA)-treated cells, responsive to this tumor-promoting agent.[3] Initial studies revealed that this activity binds to specific enhancer elements in the promoter region of the human metallothionein IIA (hMT-IIA) gene, which exhibits high basal transcription levels and responsiveness to external stimuli.[4] Concurrently, the same factor was found to interact with the enhancer of the simian virus 40 (SV40), a viral element implicated in oncogenic processes.[4] These observations established AP-1 as a key mediator of TPA-inducible gene expression through biochemical assays, such as DNase I footprinting, which demonstrated protected regions in stimulated nuclear extracts from HeLa cells.[3]The TPA-responsive element (TRE) was characterized as the consensus DNA sequence to which AP-1 binds, defined as 5’-TGAG/CTCA-3’, a palindromic motif conserved across multiple TPA-inducible promoters including those of hMT-IIA, SV40, collagenase, and stromelysin.[3] Synthetic oligonucleotides containing this TRE conferred TPA inducibility to heterologous promoters in transfection assays, confirming its functional role as an enhancer.[3] Purification of AP-1 to near homogeneity from HeLa cell nuclear extracts via sequence-specific DNA affinity chromatography yielded a 47 kDa polypeptide that specifically activated transcription from wild-type hMT-IIA templates but not from mutants lacking the TRE.[4] TPA treatment enhanced AP-1 binding activity 3- to 4-fold in nuclear extracts through a posttranslational mechanism, without altering the abundance of the factor itself.[3]These early discoveries linked AP-1 activity to the regulation of genes involved in cellular responses to tumor promoters, providing an initial connection to oncogene activation via SV40 enhancer interactions.[4] Subsequent work identified the protein components of AP-1, such as c-Fos and c-Jun, but the foundational biochemical characterization established its role as a stimulus-responsive transcription factor.[3]
Key Milestones and Oncogenic Insights
In late 1987 and early 1988, the molecular cloning of the c-jun proto-oncogene marked a pivotal advancement in understanding AP-1, as it identified the cellular homolog of v-Jun, the oncoprotein encoded by avian sarcoma virus 17 (ASV-17). Angel et al. isolated the human c-jun cDNA and demonstrated that its product functions as a sequence-specific transcriptional activator with properties akin to the previously characterized AP-1 factor, establishing a direct link between viral oncogenesis and cellular transcription regulation.[5] In December 1987, Bohmann et al. cloned the human c-jun gene and showed that it encodes a DNA-binding protein structurally and functionally related to AP-1, confirming v-Jun's role as the transforming agent in ASV-17-transformed cells and highlighting AP-1's involvement in oncogenic signaling.[6]That same year, the identification of the c-fos proto-oncogene further illuminated AP-1's oncogenic connections. Curran and Franza characterized c-Fos as the cellular counterpart to v-Fos, the oncoprotein from FBJ murine osteosarcoma virus, and proposed that Fos and Jun proteins form a heterodimeric complex responsible for AP-1 activity, thereby integrating immediate-early gene expression with tumor-inducing mechanisms.[7] This revelation underscored how retroviral oncogenes like v-fos disrupt normal cellular control by hijacking AP-1-mediated transcription.By the late 1980s and early 1990s, research established that AP-1 operates as a dimeric complex rather than a monomeric entity, expanding its regulatory scope. Halazonetis et al. demonstrated that c-Jun can homodimerize or form heterodimers with c-Fos, yielding complexes with varying DNA-binding affinities to the TPA-responsive element (TRE), a sequence motif initially identified in promoter analyses of phorbol ester-inducible genes.[8] This dimeric nature, revealed through biochemical and genetic studies, provided key insights into AP-1's versatility in transducing extracellular signals to gene expression changes implicated in oncogenesis.The 1990s brought broader recognition of AP-1's role in signal transduction via immediate-early genes, with foundational work on cellular responses to stimuli.
Molecular Structure
Protein Families and Components
The AP-1 transcription factor is primarily composed of dimeric complexes formed by proteins from the Jun and Fos families, which belong to the basic leucine zipper (bZIP) class of transcription factors.[9] The Jun family includes three main members: c-Jun, JunB, and JunD. These proteins share a high degree of structural similarity, particularly in their bZIP domains, which enable DNA binding and protein-protein interactions. All Jun family members are capable of forming homodimers among themselves or heterodimers with proteins from other families, allowing for diverse transcriptional regulation.[10] c-Jun, the prototypical member, was identified as a key component of AP-1 and exhibits potent transactivation potential, while JunB and JunD often act as modulators with weaker or context-dependent activity.[11]The Fos family consists of four proteins: c-Fos, FosB, Fra-1 (fos-related antigen 1), and Fra-2. Unlike the Jun proteins, Fos family members lack the ability to homodimerize and instead form stable heterodimers exclusively with Jun proteins to constitute functional AP-1 complexes.[9] c-Fos, the founding member, is particularly notable for its role in rapid transcriptional responses, while FosB, Fra-1, and Fra-2 contribute to sustained or specialized AP-1 activity, with Fra-1 and Fra-2 often displaying greater stability in chronic conditions.[12] These heterodimers bind to TPA-responsive elements (TREs) in DNA, facilitating gene expression control.In addition to the core Jun and Fos families, AP-1 can incorporate proteins from other bZIP families to form specialized heterodimers, expanding its functional repertoire. The ATF family, particularly ATF2 (also known as CREB-p1), can heterodimerize with Jun proteins to bind CRE/ATF sites, influencing responses to stress and growth signals.[9] Similarly, Maf family members such as c-Maf, MafB, MafA, MafG, and MafK form heterodimers with Jun or Fos to regulate tissue-specific genes, often with repressive or activating effects depending on the context.[13] The JDP family, including JDP1 and JDP2, acts predominantly as inhibitors by forming inactive heterodimers with Jun or c-Fos, thereby fine-tuning AP-1 activity.[14]Members of the Jun and Fos families are classified as immediate-early genes, characterized by their rapid transcriptional induction in response to extracellular stimuli without requiring de novo protein synthesis.[9] Their mRNAs exhibit short half-lives, typically on the order of 10-20 minutes for c-Fos and around 11 minutes for c-Jun, ensuring transient expression that allows quick adaptation to changing cellular environments.[15] Protein stability varies, with c-Fos having a half-life of approximately 2 hours and c-Jun around 90 minutes, contributing to the dynamic nature of AP-1 complexes.[16] These expression patterns underscore the role of AP-1 components in orchestrating immediate cellular responses. Dimerization occurs via the leucine zipper motif, enabling combinatorial diversity in AP-1 function.[10]
Dimerization and DNA Binding
The AP-1 transcription factor functions through dimeric complexes formed by members of the Jun, Fos, and ATF protein families, which share a conserved basic leucine zipper (bZIP) domain responsible for both dimerization and DNA binding.[17] The bZIP domain consists of two structural motifs: a basic region rich in positively charged amino acids that facilitates sequence-specific contact with DNA, and a leucine zipper region characterized by a heptad repeat of leucine residues that mediates coiled-coil dimerization between two protein monomers.[17] This architecture allows AP-1 proteins to assemble as homodimers or heterodimers, enabling versatile regulation of target genes.90147-X)Dimerization preferences among AP-1 components determine their DNA-binding specificity and efficiency. Jun family proteins, such as c-Jun, can form homodimers that bind to the palindromic 12-O-tetradecanoylphorbol-13-acetate-responsive element (TRE; consensus sequence TGAGTCA), while heterodimerization with Fos family proteins, such as c-Fos, preferentially targets the same TRE motif but with enhanced stability. In contrast, heterodimers between Jun and ATF/CREB family proteins, such as ATF-2, exhibit a preference for cyclic AMP response element (CRE)-like sites (consensus ATGACGTC), allowing AP-1 to interact with a broader repertoire of regulatory elements. These preferences arise from compatibility in the leucine zipper interfaces, where Fos lacks the ability to homodimerize effectively due to its zipper sequence, thus relying on Jun for complex formation.90147-X)Variations in DNA-binding affinity are influenced by dimer composition and the geometry of the target site. Jun/Fos heterodimers typically exhibit higher affinity for the TRE than Jun/Jun homodimers, with binding efficiencies reported up to 25-fold greater, attributed to optimal spacing of the half-sites in the 7-base pair TRE core (TGAGTCA) that aligns the basic regions for cooperative interaction.90147-X) This enhanced affinity in heterodimers stems from more stable dimerization and precise positioning on the DNA, whereas homodimers may require adjustments in half-site spacing for effective binding. For Jun/ATF heterodimers, affinity for CRE-like sites is similarly elevated compared to individual homodimers, reflecting adaptations in the basic region that accommodate the slightly divergent half-site orientation in CRE sequences.The molecular model of AP-1 DNA binding involves the leucine zipper forming a Y-shaped coiled-coil structure that positions the two basic regions to straddle the DNA helix. Upon dimerization, the basic regions extend as alpha-helices that insert into the major groove of the DNA, making direct contacts with the phosphate backbone and nucleotide bases of the TRE or CRE half-sites to achieve sequence specificity.90147-X) The zipper domain stabilizes this configuration by hydrophobic interactions between leucine residues, preventing dissociation and ensuring the fork-like grip on the DNA double helix.[17] This binding mode induces a localized conformational change in the DNA, facilitating transcriptional activation while maintaining the integrity of the dimer interface.
Evolutionary Conservation and Crystal Structures
The basic leucine zipper (bZIP) domain central to AP-1 function exhibits high evolutionary conservation, spanning from unicellular eukaryotes to mammals, with the yeast Gcn4 serving as a functional homolog that binds similar consensus sequences as AP-1 sites. This preservation underscores the domain's critical role in dimerization and DNA recognition, with sequence identity in key hydrophobic and basic residues maintained across distant taxa. Networks of bZIP interactions have diversified over a billion years while retaining core structural motifs from yeast to humans.[18][19]Orthologs of the core AP-1 components Jun and Fos are evident in key invertebrate models. In Drosophila melanogaster, the Jun ortholog DJra (also called Jra) and Fos ortholog Kayak (Kay) form heterodimers that mediate JNK signaling and developmental processes analogous to mammalian AP-1. Similarly, in Caenorhabditis elegans, jun-1 and fos-1 encode bZIP proteins that dimerize to regulate reproductive and stress responses, mirroring AP-1's DNA-binding specificity and functional partnerships. These invertebrate counterparts highlight the ancient origins of AP-1-like regulation in metazoan evolution.[20][21][22][23]Non-canonical AP-1 family members, such as those in the JDP subfamily, display more variable conservation, with homologs identified in some invertebrates like moths but absent or diverged in others, suggesting later evolutionary expansion in vertebrates.[24]High-resolution structural studies have elucidated the atomic details of AP-1 dimerization and DNA interaction. The seminal 3.0 Å crystal structure of the c-Fos/c-Jun bZIP heterodimer bound to DNA (PDB: 1FOS) revealed parallel α-helices forming a leucine zipper that clamps the DNA major groove, with basic regions inserting into adjacent half-sites of the AP-1 consensus sequence. This 1995 structure demonstrated the asymmetric interface stabilizing the heterodimer over homodimers. For the c-Jun homodimer, an NMR solution structure of the leucine zipper domain (PDB: 1JUN) illustrated the symmetrical coiled-coil arrangement, while a later 2.2 Å crystal structure of the full bZIP domain bound to CRE DNA (PDB: 1JNM) confirmed similar helical continuity and groove-spanning contacts.[25][26][27][28]Post-2000 structural analyses expanded insights into AP-1 variants and dynamics. A 3.0 Å crystal structure of the ATF2/c-Jun heterodimer bound to the interferon-β enhancer (PDB: 1T2K) highlighted interdomain contacts beyond the bZIP core, including ATF2's zinc finger motif contributing to cooperative binding with IRF-3. Complementary NMR studies of bZIP dimers, such as the c-Jun leucine zipper, revealed intrinsic flexibility in linker regions adjacent to the zipper, allowing conformational adaptability in non-bZIP activation domains during transcriptional regulation. These findings emphasize the structural plasticity enabling AP-1's diverse partnerships and responses to cellular signals.[29][30]
Activation and Regulation
Upstream Signaling Pathways
The activation of AP-1 transcription factors is primarily triggered by extracellular signals that converge on intracellular kinase cascades, particularly the mitogen-activated protein kinase (MAPK) pathways, leading to the phosphorylation and induction of AP-1 components such as c-Jun, c-Fos, and ATF2.[31] These signals include growth factors, cytokines, and stress stimuli that initiate rapid cellular responses.The c-Jun N-terminal kinase (JNK) pathway plays a central role in AP-1 activation by phosphorylating c-Jun at serine residues 63 and 73, enhancing its transcriptional activity in response to stress and inflammatory signals. JNK is activated downstream of various receptors, such as those for tumor necrosis factor-α (TNF-α), where prolonged JNK activity leads to sustained AP-1 stimulation.[32] Similarly, the extracellular signal-regulated kinase (ERK) pathway, activated by mitogenic stimuli, induces c-Fos expression through phosphorylation of the ternary complex factor Elk-1, which binds to the serum response element in the c-fos promoter.[33] The p38 MAPK pathway contributes by phosphorylating ATF2, a partner in AP-1 heterodimers, particularly in response to environmental stresses.Beyond MAPKs, other pathways influence AP-1, including protein kinase C (PKC) activation by phorbol esters like 12-O-tetradecanoylphorbol-13-acetate (TPA), which directly stimulates AP-1 binding to TPA-responsive elements. Crosstalk with NF-κB occurs in inflammatory contexts, where NF-κB can modulate JNK activity to enhance AP-1 function, amplifying proinflammatory gene expression. Growth factor receptors, such as the epidermal growth factor receptor (EGFR), activate the Ras-ERK cascade to promote AP-1-dependent proliferation signals. Cytokine receptors, exemplified by TNF-α receptors, engage JNK to drive AP-1 activation.[32]AP-1 activation exhibits temporal dynamics characteristic of immediate-early responses, with induction occurring within minutes of stimulation, enabling rapid adaptation to environmental cues before de novo protein synthesis is required. This swift onset distinguishes AP-1 from delayed responses, facilitating its role in coordinating early transcriptional events.[34]
Post-Translational Modifications
Post-translational modifications (PTMs) play a crucial role in regulating the stability, localization, and transcriptional activity of AP-1 transcription factors, enabling rapid responses to extracellular signals. These covalent changes, including phosphorylation, ubiquitination, acetylation, sumoylation, and redox modifications, modulate AP-1 dimer function without requiring de novo protein synthesis.Phosphorylation is one of the most prominent PTMs affecting AP-1, particularly targeting the N-terminal transactivation domain of c-Jun. c-Jun N-terminal kinase (JNK) phosphorylates c-Jun at serine residues 63 and 73, which enhances its transactivation potential and promotes recruitment of coactivators to target promoters. This modification is essential for AP-1 activation in stress responses, as unphosphorylated c-Jun exhibits reduced transcriptional efficacy. In contrast, glycogen synthase kinase 3β (GSK3β) phosphorylates c-Jun at threonine 239 within its DNA-binding domain, inhibiting its activity and priming it for degradation to prevent prolonged signaling.Ubiquitination controls AP-1 protein turnover via the proteasome, ensuring transient activity. The E3 ubiquitin ligase FBW7 recognizes GSK3β-phosphorylated c-Jun and promotes its polyubiquitination, leading to rapid proteasomal degradation and termination of AP-1 signaling. Similarly, c-Fos undergoes ubiquitin-mediated destabilization, with its C-terminal domain serving as a degron that facilitates proteasomal breakdown, thereby limiting the duration of inducible AP-1 complexes. Certain non-degradative ubiquitination events, such as K63-linked chains, can protect AP-1 components from proteasomal degradation under oxidative stress, stabilizing dimers for sustained activity.Acetylation and sumoylation provide additional layers of regulation by altering AP-1 interactions with chromatin and other factors. The histone acetyltransferases CBP and p300 acetylate c-Jun at lysine 271 in its DNA-binding domain, enhancing its transcriptional potency by facilitating coactivator binding and histone acetylation at target loci. Conversely, JunB is sumoylated at lysine 237, repressing its transactivation function and shifting AP-1 toward inhibitory dimers in anti-proliferative contexts.Redox modifications, driven by reactive oxygen species (ROS), directly impact AP-1 DNA binding. Oxidation of conserved cysteine residues in the basic DNA-binding region of Jun and Fos proteins forms intramolecular disulfide bonds, inhibiting TRE recognition and suppressing AP-1-dependent transcription during oxidative stress. This reversible PTM allows AP-1 to integrate redox signals, with reduction by thioredoxin restoring binding activity.
Interactions with Co-Regulators
AP-1 transcription factors interact with various co-activators to enhance chromatin accessibility and transcriptional initiation at target promoters. The co-activators CREB-binding protein (CBP) and its paralog p300 bind to the transactivation domain of c-Jun, a core AP-1 component, facilitating recruitment to AP-1 binding sites and subsequent histoneacetylation. This interaction promotes acetylation of histones H3 and H4 at promoter regions, loosening chromatin structure and enabling efficient transcription factor assembly and RNA polymerase II (Pol II) engagement. Seminal studies demonstrate that CBP/p300's histone acetyltransferase (HAT) activity is essential for AP-1-mediated activation of genes involved in cell proliferation and stress responses, with disruption of this binding impairing transcriptional output.[35][36]The Mediator complex further amplifies AP-1's transcriptional potency by bridging AP-1 dimers to the Pol II pre-initiation complex. AP-1 recruits Mediator subunits, such as MED1, to enhancers and promoters, stabilizing Pol II recruitment and phosphorylation of its C-terminal domain to transition from pausing to productive elongation. This co-activator function is particularly evident in signal-induced gene expression, where AP-1-guided Mediator assembly coordinates with general transcription factors to boost Pol II processivity at immediate-early genes like c-fos.[37][38]In contrast, co-repressors such as nuclear receptor corepressor 1 (NCoR1) and silencing mediator for retinoid and thyroid hormone receptors (SMRT) suppress AP-1 activity through recruitment of histone deacetylases (HDACs). NCoR1/SMRT complexes bind unphosphorylated c-Jun at AP-1 sites, docking HDAC3 to deacetylate histones and condense chromatin, thereby silencing inflammatory gene networks in macrophages and other immune cells. This HDAC-mediated repression integrates anti-inflammatory signals, preventing excessive AP-1-driven proinflammatory responses during homeostasis.[39][40]The Groucho/TLE family of co-repressors, including TLE1, modulates AP-1 output by interacting with JunB to attenuate transcription at select promoters. TLE proteins are recruited via short motifs in JunB, recruiting HDACs and promoting chromatin compaction to repress genes involved in differentiation and cell cycle control. This repression is context-dependent, often counterbalancing activating AP-1 dimers in developmental and oncogenic settings.[41]Chromatin remodeling complexes like SWI/SNF also cooperate with AP-1 to overcome nucleosomal barriers. As a pioneer factor, AP-1 binds closed chromatin and recruits the BAF variant of SWI/SNF via interactions with subunits like ARID1A and SMARCC1, driving ATP-dependent nucleosome eviction and enhanced DNA accessibility at enhancers. This facilitates AP-1's access to embedded binding sites, reshaping 3D chromatin landscapes for sustained transcriptional activation during epithelial-mesenchymal transitions and stress adaptation. Over 90% of AP-1 peaks colocalize with SWI/SNF occupancy, underscoring their interdependent roles in dynamic gene regulation.[42][43]AP-1 engages in crosstalk with other regulators to fine-tune responses to cellular cues. In oxidative stress, AP-1 cooperates with Nrf2 at composite promoter elements, such as in the sulfiredoxin gene (Srx), where both factors synergistically drive expression of antioxidant enzymes to mitigate reactive oxygen species damage. This interaction enhances cytoprotection without competing for binding sites, integrating AP-1's proliferative signals with Nrf2's detoxifying program.[44]Similarly, AP-1 intersects with β-catenin in Wnt signaling, where c-Jun physically associates with β-catenin/TCF complexes at promoters, amplifying transcription of shared targets like c-Myc and cyclin D1 to promote cell proliferation and tumorigenesis. This crosstalk occurs via direct protein-protein contacts, enabling Wnt to potentiate AP-1 activity in intestinal and cancer contexts, though it can also lead to mutual antagonism depending on signaling context. Dimer composition influences co-regulator preference, with Fos-Jun heterodimers favoring CBP/p300 over repressors like NCoR.[45][46]
Biological Functions
Cell Proliferation and Senescence
The AP-1 transcription factor complex, particularly dimers composed of c-Jun and Fos family members, plays a pivotal role in promoting cell proliferation by directly inducing the expression of key cell cycle regulators such as cyclin D1, cyclin A, and cyclin-dependent kinases (CDKs). These targets facilitate the G1-to-S phase transition, enabling progression through the cell cycle in response to mitogenic signals. For instance, c-Jun/Fos heterodimers bind to AP-1 sites in the cyclin D1 promoter, driving its transcription and subsequent activation of CDK4/6 complexes that phosphorylate the retinoblastoma protein, thereby releasing E2F transcription factors to promote S-phase entry.[47][48][49]In contrast, certain AP-1 subunits exert inhibitory effects on proliferation; notably, JunB acts as a negative regulator by repressing c-Jun-dependent targets, including through direct transcriptional activation of the cyclin-dependent kinase inhibitor p16^INK4a, which enforces G1 arrest. This antagonistic function of JunB highlights the context-dependent nature of AP-1 activity in balancing proliferative signals.[50][51]AP-1 also contributes to cellular senescence, a state of permanent cell cycle arrest, particularly in oncogene-induced senescence (OIS). Fra-1/JunD dimers are key mediators, upregulating senescence effectors like p16^INK4a and p21^CIP1 to impose G1 arrest and remodel the chromatin landscape at senescence-associated enhancers. These dimers pioneer enhancer accessibility, facilitating the expression of senescence-associated secretory phenotype (SASP) factors such as IL1A and IL1B, which reinforce the senescent state. Studies in RAS-driven OIS models demonstrate that perturbing AP-1 activity, including Fra-1 and JunD, can reverse this transcriptional program, underscoring its reversible nature.[52][53]Experimental evidence from mouse models supports these roles; for example, conditional inactivation of AP-1 factors in the epidermis leads to basal keratinocyte hyperproliferation and disrupted cell cycle control, manifesting as thickened skin layers and delayed differentiation.[54]
Differentiation and Development
The AP-1 transcription factor complex, particularly involving c-Jun and c-Fos, plays a pivotal role in keratinocyte terminal differentiation by activating the promoters of differentiation-specific markers such as keratins K1 and K10. In primary human keratinocytes, overexpression of c-Jun and c-Fos transactivates the K1 promoter through AP-1 binding sites, while c-Jun alone enhances K10 promoter activity, thereby promoting the expression of these suprabasal keratins essential for epidermal barrier formation.[55] This activation occurs in response to differentiation cues like elevated calcium levels, where AP-1 integrates signals from the JNK pathway to drive commitment to the differentiated state. Conversely, disruption of AP-1 function, such as through dominant-negative c-Jun (TAM67) expression in suprabasal epidermis, impairs the upregulation of late differentiation genes, underscoring AP-1's necessity for completing the keratinization program.[56]In bone remodeling, AP-1 dimers containing Fra-2 and JunD maintain the balance between osteoblast and osteoclast activities by positively regulating osteoblast differentiation and extracellular matrix production. Conditional knockout of Fra-2 in osteoblast lineage cells of mice results in severe osteopenia due to defective osteoblast maturation, reduced collagen type I synthesis, and impaired mineralization, highlighting Fra-2/AP-1 as a critical driver of bone formation.[57] Similarly, JunD promotes osteoblast differentiation by antagonizing inhibitory factors like menin, with JunD-deficient mice exhibiting enhanced osteoblast activity and increased bone mass, demonstrating its role in fine-tuning osteoblast-osteoclast coupling.[58][59] These findings from genetically modified mice reveal AP-1's context-dependent regulation of bone cell fate, where specific dimers ensure coordinated remodeling without overlapping proliferative effects.During mouse embryonic development, AP-1 components such as c-Jun and JunB are essential for neural crestcell migration and limb bud formation, often integrating BMP signaling to pattern tissues. In c-Jun knockout embryos, defects in neural crest-derived structures arise from impaired cell migration, as c-Jun mediates JNK-dependent delamination and motility in response to BMP gradients along the neural tube.[60] while in limb buds, c-Jun-containing AP-1 dimers regulate interdigital cell death and patterning via BMP-responsive elements, ensuring proper anterior-posterior axis formation.[61] These roles position AP-1 as a downstream effector of BMP signaling in orchestrating embryonic cell fate decisions and morphogenetic movements.AP-1 family members, including c-Fos, are crucial for placental trophoblast invasion and uterine implantation in mice, facilitating the maternal-fetal interface. Although c-Fos knockout mice are viable, expression of c-Fos in trophoblast cells supports invasive behavior during early placentation, with its induction correlating with implantation window preparation in the endometrium.[62] More prominently, related AP-1 proteins like Fra-1 (Fosl1) are indispensable, as Fosl1-null embryos fail to undergo proper trophoblast invasion into uterine spiral arteries, resulting in defective vascular remodeling and mid-gestational lethality due to impaired implantation and nutrient exchange.[63] Dimer-specific activities, such as c-Fos/c-Jun complexes, further modulate trophoblast motility by regulating matrix metalloproteinase expression essential for decidual penetration.[64]
Apoptosis and Stress Response
The AP-1 transcription factor plays a dual role in apoptosis, promoting programmed cell death in certain contexts while inhibiting it in others, depending on the cellular environment and dimer composition. In pro-apoptotic scenarios, c-Jun, a core component of AP-1, drives the expression of Fas ligand (FasL), which activates the extrinsic apoptosis pathway following ultraviolet (UV) irradiation. This induction occurs through JNK-mediated phosphorylation of c-Jun, enabling AP-1 binding to the FasL promoter and triggering caspase-dependent cell death in keratinocytes and fibroblasts. Similarly, c-Jun upregulates the BH3-only protein Noxa, a p53 target that sensitizes mitochondria to pro-apoptotic signals, thereby enhancing UV-induced apoptosis in human skin cells.[65] These mechanisms highlight c-Jun's role in eliminating damaged cells to prevent mutagenesis.The JNK-AP-1 axis is particularly critical in neuronal apoptosis, where sustained JNK activation phosphorylates c-Jun, leading to transcriptional induction of pro-death genes. In sympathetic neurons deprived of nerve growth factor, c-Jun deficiency blocks apoptosis, demonstrating its essential function downstream of JNK in cytochrome c release and caspase activation.[66] This pathway operates independently of p53 in some neuronal models, underscoring AP-1's direct contribution to developmental and stress-induced neuronal elimination.[67]Conversely, AP-1 can exert anti-apoptotic effects through specific dimers, such as Fra-1-containing complexes that upregulate anti-apoptotic Bcl-2 family members. In transformed cancer cells, Fra-1 binds to the Bcl-XL promoter, elevating its expression and inhibiting mitochondrial outer membrane permeabilization, thereby promoting cell survival under genotoxic stress.[68] This protective function is evident in breast and lung cancer models, where Fra-1 overexpression correlates with resistance to chemotherapy-induced death.[69]In the stress response, AP-1 dimers involving ATF2 and Jun proteins coordinate adaptation to DNA damage by synergizing with p53. The Jun/ATF2 heterodimer binds to AP-1 sites in the promoters of DNA repair genes like GADD153, enhancing p53-dependent transcription and facilitating cell cycle arrest or repair rather than immediate apoptosis.[70] This synergy is activated by ATM/ATR kinases following genotoxic insults, allowing cells to prioritize survival.[71] Additionally, c-Fos contributes to heat shock responses by inducing heat shock protein 70 (HSP70), which chaperones misfolded proteins and prevents aggregation during thermal stress.[72]The context-dependence of AP-1 in apoptosis and stress arises from dimer switching, where JunD-containing complexes provide protection against oxidative stress. Unlike pro-apoptotic c-Jun homodimers, JunD/Fos heterodimers upregulate antioxidant genes such as MnSOD, reducing reactive oxygen species (ROS) levels and preventing endothelial and podocyte damage in models of vascular injury.[73] This protective role is further supported by post-translational modifications, such as phosphorylation, that modulate ROS-sensitive dimerization.00804-9)
Physiological Roles
Tissue-Specific Regulation
In the liver, AP-1 activity, particularly through dimers involving JunB and Fra-1, contributes to hepatocyte regeneration following injury such as partial hepatectomy or toxin exposure. These dimers are induced as part of the immediate-early gene response, supporting cell proliferation and survival during the regenerative process by modulating genes involved in detoxification and anti-apoptotic pathways. For instance, Fra-1 overexpression enhances protection against acetaminophen-induced injury by upregulating glutathione S-transferase P1 (GSTP1) in hepatocytes, thereby facilitating recovery.[74][75]In the skin, c-Jun plays a pivotal role in AP-1-mediated responses to injury and environmental stress, driving keratinocyte migration and proliferation essential for wound healing and UV-induced repair. During wound healing, c-Jun promotes re-epithelialization by regulating genes that facilitate epithelial cell movement across the wound bed, and its conditional knockout in keratinocytes results in delayed closure and impaired re-epithelialization due to defective migration. Similarly, UV irradiation rapidly activates c-Jun, leading to AP-1 dimer formation that initiates protective responses like matrix remodeling and inflammation resolution in epidermal cells.[76][77][78]In the brain, AP-1 complexes formed by ATF3 and Jun family members, such as c-Jun, are critical for neuronal survival in response to ischemic injury. Following cerebral ischemia, ATF3 and phosphorylated c-Jun are upregulated in neurons, where they cooperatively induce heat shock protein 27 (Hsp27) expression, promoting anti-apoptotic mechanisms and axon regeneration to mitigate cell death. ATF3 deficiency exacerbates neuronal apoptosis post-ischemia, underscoring its protective function in this context via interaction with c-Jun pathways.[79][80]In the heart, c-Fos activation within AP-1 dimers is integral to pathological hypertrophy triggered by angiotensin II signaling. Angiotensin II rapidly induces c-fos expression in cardiac myocytes through G-protein-coupled receptor pathways, leading to AP-1 binding that drives hypertrophic gene programs and cell enlargement. This response contributes to adaptive remodeling under hemodynamic stress, with c-Fos serving as a key immediate-early mediator in the transition to hypertrophy.[81][82]Tissue-specific regulation of AP-1 often involves preferential dimer compositions, such as JunB/Fra-1 in the liver or c-Jun/ATF3 in the brain, reflecting local isoform expression patterns that fine-tune responses to injury.[74]
Role in Immune Response and Inflammation
AP-1 transcription factors, particularly the c-Jun/c-Fos heterodimers, play a central role in cytokine production during T-cell activation. Upon T-cell receptor (TCR) and CD28 co-stimulation, these dimers bind to TPA-responsive elements (TREs) in the promoters of interleukin-2 (IL-2) and tumor necrosis factor-alpha (TNF-α), driving their transcription.[83][84] Specifically, c-Jun phosphorylation by JNK enhances AP-1 binding to composite NFAT/AP-1 sites in the IL-2 promoter, essential for T-cell proliferation and effector function.[85] Similarly, AP-1 activation via ERK and JNK pathways induces TNF-α expression in T-cells, contributing to inflammatory signaling.[86] This regulation underscores AP-1's integration with upstream signaling, such as the JNK pathway in Toll-like receptor (TLR) responses, to orchestrate adaptive immune activation.[85]In macrophages, AP-1 components like Fra-1 promote pro-inflammatory activation and polarization. Fra-1 binds to the IL-6 promoter in response to stimuli such as lipopolysaccharide (LPS), enhancing IL-6 production that acts autocrine to drive macrophage skewing toward inflammatory phenotypes.[87] Although primarily linked to alternative activation in some contexts, Fra-1 contributes to the expression of inducible nitric oxide synthase (iNOS) and other M1-associated genes by facilitating AP-1-mediated transcription of pro-inflammatory mediators during innate immune responses.[88] This mechanism amplifies nitric oxide production and pathogen clearance but can perpetuate chronicinflammation if dysregulated.[89]AP-1 hyperactivity is implicated in autoimmune conditions, notably rheumatoid arthritis (RA), where elevated c-Jun activity in synovial fibroblasts sustains inflammation. Nuclear extracts from RA synovial tissues exhibit significantly higher AP-1 DNA-binding compared to osteoarthritis controls, correlating with increased expression of matrix metalloproteinases and cytokines that drive joint destruction.[90] The AP-1 family member JunD promotes inflammatory activation in macrophages. In JunD-deficient macrophages, pro-inflammatory gene expression—including TNF-α and IL-6—is reduced, leading to dampened activation and cytokine secretion.[91] This function of JunD supports macrophage responses in immune homeostasis.
Involvement in Cancer and Other Diseases
The AP-1 transcription factor exhibits paradoxical roles in cancer, acting as both an oncogene and tumor suppressor depending on the cellular context and family member involved. In lung cancer, particularly non-small cell lung cancer (NSCLC), c-Jun promotes tumorigenesis through activation of downstream pathways that enhance cell survival and resistance to therapies such as osimertinib.[92] Overexpression of c-Jun has been linked to aggressive phenotypes in NSCLC, where it drives metabolic reprogramming and glutaminase expression to support tumor growth.[93] Similarly, in breast cancer, Fra-1 (encoded by FOSL1) facilitates invasion by inducing matrix metalloproteinase 9 (MMP9) expression, enabling extracellular matrix degradation and metastasis; elevated Fra-1 levels correlate with poor prognosis in triple-negative breast cancer subtypes.[94][95]Conversely, certain AP-1 components exert tumor-suppressive effects. Loss or downregulation of JunB in B-lymphoid cells promotes proliferation and transformation, contributing to lymphoma development, as evidenced by surveys of human lymphoma samples showing reduced JunB expression in aggressive cases.[51] In anaplastic large cell lymphoma (ALCL), deletion of JunB alongside c-Jun exacerbates tumor progression by altering signaling pathways like PDGFRβ.[96] These findings highlight JunB's role in restraining oncogenic signaling, including cyclin induction during proliferation.[97]Beyond cancer, AP-1 dysregulation contributes to non-oncologic diseases. In psoriasis, activation of c-Jun/AP-1 in keratinocytes drives hyperproliferation and inflammatory cytokine production, exacerbating skin lesions; Fra-1 overexpression further shifts keratinocytes toward an epithelial-mesenchymal transition-like state.[98][99] In neurodegeneration, such as amyotrophic lateral sclerosis (ALS), upregulated c-Jun in motor neurons promotes axonal degeneration and autophagic responses, with phosphorylated c-Jun observed in affected spinal cord tissues.[100][101]Therapeutic strategies targeting AP-1 have advanced, particularly for disease-associated dysregulation. The selective AP-1 inhibitor T-5224, which blocks c-Fos/c-Jun dimers, has shown preclinical efficacy in reducing fibrosis by suppressing proinflammatory gene expression and is under investigation for inflammatory conditions like pulmonary fibrosis.[102] Post-2020 developments include proteolysis-targeting chimeras (PROTACs) designed to degrade AP-1 family members; for instance, T-5224-based PROTACs selectively degrade Fra-1 (FOSL1) in head and neck squamous cell carcinoma, reducing cancer stemness without broad toxicity.[103] These approaches underscore AP-1's potential as a druggable target in cancer and fibrosis.
Target Genes and Regulome
Consensus Binding Sites and Motifs
The AP-1 transcription factor, composed of bZIP domain-containing proteins such as Jun and Fos family members, primarily recognizes palindromic DNA sequences known as TPA-responsive elements (TREs). The canonical TRE consensus sequence is 5'-TGAGTCA-3', a 7-base pair motif identified in the promoters of phorbol ester-inducible genes, including the human interstitial collagenase (MMP-1) gene. This sequence allows for symmetric binding by AP-1 dimers, with each monomer contacting a half-site centered on the invariant TGAC core.Variants of the consensusmotif exist, influenced by the specific dimer composition. For instance, Jun/ATF heterodimers preferentially bind to CRE-like sequences, such as 5'-ATGACGTCAT-3', which features an additional central GC pair compared to the TRE, enabling recognition by ATF/CREB family proteins in complex with Jun. In contrast, Fos/Jun heterodimers exhibit higher affinity for the strict TRE motif, while Jun homodimers show broader tolerance but reduced efficiency on TRE compared to CRE sites. Flanking sequences adjacent to the core motif modulate binding affinity; for example, GC-rich contexts enhance specificity and stability for Fos/Jun dimers by influencing DNA shape and bendability.Non-consensus AP-1 sites, which deviate from the ideal TRE, can still support dimer binding but often with altered specificity. In the promoter of the TIMP-1 gene, a variant sequence 5'-TGAGTAA-3' serves as a functional AP-1 site, where the terminal A substitution reduces affinity for classical Fos/Jun but accommodates other dimers under specific signaling conditions.[104] In vitro selection methods, such as SELEX, have confirmed these preferences: Fos/Jun dimers enrich for TRE motifs with 7-bp spacing, while Jun/ATF combinations favor 8-bp spaced CRE variants, highlighting how dimer composition dictates motif selection.
Genome-Wide Target Identification
Genome-wide identification of AP-1 binding sites has primarily relied on chromatin immunoprecipitation followed by sequencing (ChIP-seq), which maps the locations of AP-1 subunits across the genome in various cell types and conditions. In stimulated mouse embryonic fibroblasts (MEFs), ChIP-seq analyses of AP-1 family members such as FOS, FOSL2, and JUND have identified approximately 55,919 binding sites, with over 90% located in promoter-distal regions including enhancers. These sites are enriched near active regulatory elements, such as those marked by H3K4me1 and accessible chromatin, highlighting AP-1's role in signal-dependent enhancer activation.[105]Cell-type specificity is evident in immune cells, where lipopolysaccharide (LPS) stimulation of bone marrow-derived macrophages (BMDMs) induces dynamic AP-1 binding. For instance, ChIP-seq for JunD in LPS-treated rat BMDMs revealed 18,124 binding peaks associated with 7,612 genes, predominantly at distal enhancers responsive to inflammatory signals.[106] In mouse macrophages stimulated with Kdo2-lipid A (a TLR4 agonist mimicking LPS), over 10,000 sites were detected for subunits like ATF3, Jun, and JunD, with additional inducible binding for Fos and JunB, totaling more than 50,000 unique AP-1 sites across family members.[107] These findings underscore AP-1's context-dependent occupancy, with stimulation expanding binding to thousands of inflammatory response elements.In cancer cells, AP-1 binding exhibits similar scale but with distinct regulatory features. ChIP-seq in triple-negative breast cancer (TNBC) BT549 cells identified 11,670 Fra-1 sites and 14,201 c-Jun sites, with substantial overlap (84%) and enrichment at enhancers, including those within super-enhancers like the ZEB2 locus that drive invasiveness. Broader analyses in lung adenocarcinoma (A549) and leukemia (K562) cells confirm AP-1 hotspots—multi-subunit binding regions numbering 1,000 to 4,000—often overlapping super-enhancers, though only about 15% colocalize precisely, emphasizing AP-1's contribution to oncogenic enhancer networks.Advances since 2010 have integrated ChIP-seq with RNA-seq to link AP-1 binding to transcriptional output. In stimulated macrophages and fibroblasts, such integrations demonstrate that AP-1 occupies 20-30% of stimulus-inducible genes, particularly those involved in inflammation and stress responses, with binding correlating to dynamic gene expression changes. For example, in LPS-stimulated BMDMs, JunD ChIP-seq peaks overlap with differentially expressed genes, confirming direct regulation of oxidative stress and cytokine pathways. These multi-omics approaches reveal AP-1's selective activation of subsets of bound sites based on cellular context.[106]Public databases like ENCODE provide comprehensive AP-1 data, linking its binding to open chromatin regions across cell types. ENCODE ChIP-seq datasets show AP-1 motifs enriched in accessible chromatin (e.g., DNase I hypersensitive sites) in over 100 humancell lines, with AP-1 occupancy facilitating co-binding of other factors at enhancers. In oesophageal adenocarcinoma models using ENCODE-referenced open chromatin profiles, AP-1 motifs appear in 65% of differentially accessible regions, associating with upregulated oncogenic genes. These resources enable cross-cell comparisons, revealing AP-1's conserved role in chromatin remodeling and gene regulation.[108]
Functional Networks and Context-Dependence
AP-1 integrates into core regulatory networks that drive key cellular processes, particularly proliferation, inflammation, and metastasis. In proliferation networks, AP-1 activates genes such as cyclin D1 and c-Myc, which promote cell cycle progression and oncogenic growth; for instance, AP-1 dimers bind to the cyclin D1 promoter in an AKT-dependent manner to enhance transcription during mitogenic signaling.[109] Similarly, in inflammatory circuits, AP-1 upregulates COX-2 and IL-6, amplifying pro-inflammatory responses; AP-1 binds directly to the COX-2 promoter upstream of the transcription start site to induce expression in response to stimuli like TNFα.[110] For metastasis, AP-1 targets matrix metalloproteinases (MMPs) and vascular endothelial growth factor (VEGF), facilitating extracellular matrix remodeling and angiogenesis; AP-1 mediates VEGF-induced endothelial cell migration and proliferation by activating downstream transcriptional programs.[111] These networks highlight AP-1's role as a hub coordinating multiple oncogenic and inflammatory pathways.The functional output of AP-1 is highly context-dependent, shaped by the specific dimer composition and co-regulators present in a given cellular environment. Different Jun/Fos heterodimers can either activate or repress target genes; for example, c-Jun/c-Fos dimers typically promote transcription, while JunD-containing complexes often exert repressive effects, such as in quiescent cells where JunD inhibits proliferation-associated genes.[112] Co-regulators further modulate this, with factors like p300/CBP enhancing activation in proliferative contexts or HDACs promoting repression during stress; this dimer-specific and co-regulator-driven variability allows AP-1 to toggle between pro- and anti-tumorigenic roles depending on signals like MAPK pathway intensity.[113]AP-1 participates in intricate feedback loops that fine-tune its own activity and amplify signaling cascades. It auto-regulates the jun and fos genes through direct binding to their promoters, creating positive feedback that sustains AP-1 levels during prolonged stimulation; the c-jun promoter, in particular, is positively autoregulated by Jun/AP-1 to reinforce immediate-early gene expression.[114] Additionally, AP-1 engages in crosstalk with other transcription factors, such as STAT3, within inflammatory networks; in cytokine storms, AP-1 and STAT3 co-activate shared targets like IL-6, exacerbating systemic inflammation through a cooperative NF-κB-linked circuit.[115]Recent single-cell RNA sequencing studies have illuminated AP-1's role as a central hub in tumor microenvironments, revealing heterogeneous activation patterns across cell types. Post-2020 analyses in clear cell renal cell carcinoma show AP-1 coordinating immune evasion and stromal remodeling at single-cell resolution, with elevated Jun/Fos activity in tumor-associated macrophages driving pro-metastatic inflammation.[116] In glioblastoma, scRNA-seq identifies AP-1 as a plasticity regulator, where its inhibition disrupts tumor cell states and microenvironmental interactions, highlighting context-specific hubs that vary by tumor subtype and therapy response.[117] These insights underscore AP-1's dynamic integration into multicellular networks, beyond bulk analyses.