Abstract
Actin and microtubules play essential roles in aberrant cell processes that define and converge in cancer including: signaling, morphology, motility, and division. Actin and microtubules do not directly interact, however shared regulators coordinate these polymers. While many of the individual proteins important for regulating and choreographing actin and microtubule behaviors have been identified, the way these molecules collaborate or fail in normal or disease contexts is not fully understood. Decades of research focus on Profilin as a signaling molecule, lipid-binding protein, and canonical regulator of actin assembly. Recent reports demonstrate that Profilin also regulates microtubule dynamics and polymerization. Thus, Profilin can coordinate both actin and microtubule polymer systems. Here we reconsider the biochemical and cellular roles for Profilin with a focus on the essential cytoskeletal-based cell processes that go awry in cancer. We also explore how the use of model organisms has helped to elucidate mechanisms that underlie the regulatory essence of Profilin in vivo and in the context of disease.
Keywords: Profilin, Profilin-1, Profilin-2, Actin, Microtubules, Cytoskeletal crosstalk, Motility, Cell division
1. INTRODUCTION: CELLULAR INFRASTRUCTURE CONSTRUCTED FROM CYTOSKELETAL BUILDING BLOCKS
Actin and microtubules bestow useful properties to cells including: paths for transport, shape, and the forces that propel and steer movement (Fig. 1). The actin and microtubule cytoskeletons are traditionally studied as separate entities, restricted to specific cellular regions, and regulated by different suites of binding proteins unique for each polymer. Mounting genetic and pharmacological evidence suggests a rich degree of functional crosstalk and physical interactions between both systems, ultimately manifesting in the normal function of fundamental cellular processes, e.g., motility, adhesion, intracellular transport, wound healing, phagocytosis, and cell division (Coles and Bradke, 2015; Dogterom and Koenderink, 2019; Etienne-Manneville, 2004; Li and Gundersen, 2008; Rodriguez et al., 2003). Many proteins likely to be involved in actin-microtubule interactions have already been identified and characterized with regard to either actin or microtubules alone, owing to pioneering work in genetically tractable model organisms (Chang and Martin, 2009; Coles and Bradke, 2015; Dogterom and Koenderink, 2019; Etienne-Manneville, 2004; Gardiner and Marc, 2011; Prokop et al., 2013; Roeles and Tsiavaliaris, 2019; Slater et al., 2017). However, the detailed mechanisms that underlie cytoskeletal synergy and that go wrong in cancer and disease are just starting be investigated (Akhmanova and Steinmetz, 2010; Coles and Bradke, 2015; Dogterom and Koenderink, 2019; Mitchison and Kirschner, 1984a, 1984b; Rodriguez et al., 2003; Salmon et al., 2002). Here we focus on Profilin as a regulator of actin and microtubules in biochemical assays, and cancer-relevant cell processes and signaling pathways.
Fig. 1.
Cellular actin and microtubule architectures. Cartoon of actin (green) and microtubule (purple) structures in a motile cell. (A) Diagram of actin contact with focal adhesion proteins including: Integrins (green), Vinculin (purple), Talin (red), and Paxillin (blue). Actin filaments are bundled by α-actinin (magenta). (B) Leading edge lamellipodial actin branched actin structures are generated by the Arp2/3 complex. (C) Formins (navy) and Ena/VASP (yellow) elongate straight filaments bundled by Fascin proteins (cyan) in filopodial structures. Inset: Diagram of nucleation steps for actin and microtubules. NPF, nucleation promoting factor. +, the faster growing end of actin or microtubules. −, the slower growing end of actin or microtubules.
Actin filaments and microtubules share many properties. All cells contain both polymers in subunit-based and polarized filament forms (Fig. 1, inset). New assembly of either polymer is kinetically unfavorable and requires a template to organize polymer growth. For actin: de novo assembly requires a seed of three to four monomers that can further polymerize into helical two-stranded filaments with a diameter ~ 7 nm (Barshop et al., 1983; Cooper et al., 1983; Courtemanche, 2018; Oda et al., 2016; Sept and McCammon, 2001; Skruber et al., 2018; Svitkina et al., 1997). For microtubules: spontaneous microtubule assembly can occur in vitro with high concentrations of tubulin and time (Caudron et al., 2000; Desai and Mitchison, 1997; Fygenson et al., 1995; Job et al., 2003; Nogales, 2001; Roostalu and Surrey, 2017). In cells the critical concentration required for microtubule assembly far exceeds the amount of available free tubulin (estimated ~ 20 μM) (Fygenson et al., 1995; Voter and Erickson, 1984; Wieczorek et al., 2015). Thus, cellular microtubule assembly requires a stable template (i.e., a ring-shaped complex that mimics microtubule dimensions or a severed microtubule) to ultimately form a ~ 25 nm cylinder held together through the lateral contact of ~ 13 parallel protofilaments (this number varies depending on cell source and experiment) (Chaaban and Brouhard, 2017; Kollman et al., 2010; Roostalu and Surrey, 2017; Wieczorek et al., 2015). Protofilaments are formed from αβ-tubulin heterodimers that intrinsically self-assemble in a head-to-tail fashion and impart structural polarity to the microtubule, with an exposed β-subunit pointing outward (Alushin et al., 2014; Löwe et al., 2001; Mitchison, 1993; Nogales, 2001; Nogales et al., 1998). Cells employ a plethora of actin and microtubule nucleation promoting factors to overcome these barriers to nucleation at specific times and locations (for actin: the Arp2/3 complex, Formins, Ena/Vasp, Spire; for microtubules: ɣ-Turc, XMAP215) (Chesarone et al., 2010; Courtemanche, 2018; Kollman et al., 2010; Kovar et al., 2006; Krause et al., 2003; Machesky et al., 1999; Moritz et al., 2000; Mullins and Pollard, 1999; Mullins et al., 1998; Oakley et al., 2015; Pollard, 2007; Popov et al., 2002; Pruyne et al., 2002; Quinlan et al., 2005; Sagot et al., 2002; Thawani et al., 2018). Shared actin and microtubule nucleation proteins can also link dynamic cytoskeletal behaviors in cells and reconstituted systems (Chang, 2000; Colin et al., 2018; Elie et al., 2015; Gaillard et al., 2011; Henty-Ridilla et al., 2016; Inoue et al., 2019; Kita et al., 2019; Lewkowicz et al., 2008; Plessner et al., 2019; Prezel et al., 2018; Szikora et al., 2017). For example, Profilin can block actin filament assembly rendering building blocks sterically or nucleotide assembly incompetent (De La Cruz et al., 2000; Skruber et al., 2018) and it can also stabilize parameters of microtubule growth at concentrations below the normal critical concentration required for assembly (Henty-Ridilla et al., 2017).
Many cellular actin and microtubule structures like filopodia or the mitotic spindle apparatus are not long lived. Actin and microtubules each display periods of growth (although on different time scales) and rapid disassembly. Dismantling either cytoskeletal polymer is part of a recycling process required for polymer growth to occur again, elsewhere. Many disassembly factors recognize the nucleotide state of actin (ATP, ADP-Pi, or ADP) or microtubules (GTP, GTP-Pi, or GDP) for their function, which can result in a targeted destruction of “aged” regions of polymer (Berro et al., 2010; Brieher, 2013; Manandhar et al., 2018; Margolis, 1981; Pollard and Borisy, 2003). The role of actin or microtubule disassembly factors in coordinating both cytoskeletal polymers have not been extensively studied. However, Profilin promotes actin filament disassembly by sequestering actin monomers and sterically blocking actin assembly (Carlier et al., 1993; Pantaloni and Carlier, 1993). Compared to actin filaments, microtubules have an additional intrinsic disassembly/recycling property (i.e., dynamic instability) that results in the co-existence of growing and shrinking microtubules (Burbank and Mitchison, 2006; Erickson and O’Brien, 1992; Mitchison and Kirschner, 1984a,b; Zhang et al., 2015). Catastrophe events, denoted by the stochastic switch from periods of microtubule growth to rapid depolymerization, may or may not be rescued. Thus, dynamic instability may provide an example of how physical linkages between actin and microtubules disengage. In microtubule dynamics assays Profilin did not have a significant effect on these parameters of microtubule disassembly (Henty-Ridilla et al., 2017). However, because many parameters of actin and microtubule dynamics are concentration dependent, control of either subunit pool through assembly-disassembly dynamics may ultimately influence the coordinated behaviors manifested during diverse cell processes.
In cancers, the unchecked dynamics (nucleation, assembly, disassembly, and crosstalk) of either actin or microtubules results in abnormal cell division, aberrant cell morphology or size, and invasive cell migration (Bae et al., 2009; Ding et al., 2014; Hall, 2009; Mouneimne et al., 2012; Roy and Jacobson, 2004; Tang et al., 2015). Many tumor suppressors (APC, N- and E-cadherin, neurofibromin) interact with both microtubules and actin and have been proposed to disrupt epithelial-mesenchymal transitions (EMT), epithelial positioning, and many aspects of cell migration and division (Hernandez and Tirnauer, 2010; Juanes et al., 2017, 2019). Thus, Profilin is posed as a convergence point linking the cytoskeleton and the signaling pathways that regulate cancer onset, progression, and severity in several tissues. Below we explore the biochemical and cellular facets of Profilin in coordinating actin and microtubules in normal cells, cancer, and lessons learned from unique model organisms.
2. THE PROFILIN ORIGIN STORY
Profilin was first isolated and crystalized from calf spleen extracts as an unrecognized protein present in a 1:1 complex with monomeric actin (Fig. 2A) (Carlsson et al., 1976a). This low molecular weight contaminant was later characterized as the first actin monomer binding protein—remarkable because the hypothesis that a large pool of unpolymerized actin could exist in muscle or non-muscle cells was an emerging idea at the time (Carlsson et al., 1976b, 1977; Tilney, 1976). The association of this unknown protein was sufficient to fully explain the presence of a monomeric actin state in various cell extracts (Carlsson et al., 1977). Profilin was proposed to be a rapidly reversible storage solution, as sequestered actin monomers could be rapidly interconverted between different polymerization states to promote new actin filament formation to adapt to changing cell requirements (Carlsson et al., 1976b; Tilney, 1976). To date, no eukaryotic cells have been described that do not contain actin or Profilin, and both proteins are among the most abundant proteins on Earth (Beck et al., 2011; Dominguez and Holmes, 2011).
Fig. 2.
Structural Features of Profilin. (A) View of Profilin-1 (purple) interacting with actin (gray) and with the poly-L-proline region of VASP (yellow). Important Profilin residues for interactions with actin (blue), microtubules (teal), and poly-L-proline (PLP) peptides (orange) are highlighted. Structures modeled using PDBIDs 2PAVE and 2BTF (Ferron et al., 2007; Schutt et al., 1993). (B) Profilin-1 (purple) and Profilin-2 (green) structures are very similar (RMSD = 0.3), as evident from the aligned structures. Structures modeled using PDBIDs 1FIK and 1D1J (Nodelman et al., 1999). (C) Table of amino acid similarity from BLASTp alignments for Human Profilin isoforms.
Profilins are small (12–15 kDa), well conserved, and highly abundant (Carlsson et al., 1976b; Pollard et al., 2000; Witke, 2004; Witke et al., 1998). They are essential in all forms of life from Archaea to eukaryotes, and even some viruses (Akıl and Robinson, 2018; Zaremba-Niedzwiedzka et al., 2017). Profilins exist as a single gene in many organisms (yeast, Drosophila, Acanthamoeba, certain viruses, Toxoplasma) and as several isoforms in others (Arabidopsis, Chlamydomonas, humans, worms, Dictyostelium). The greatest number and diversity of Profilin isoforms comes from Maize (10; and 2 additional non-annotated sequences); however, more diversity may be possible in higher ploidy phytozome genomes (Bao et al., 2011). The role of Profilin as a major regulator of actin assembly is broadly conserved in each of these systems (Akıl and Robinson, 2018; Di Nardo et al., 2000; Dominguez and Holmes, 2011; Witke, 2004; Zaremba-Niedzwiedzka et al., 2017). Most Profilins have highly conserved actin-, poly-L-proline (PLP)-, and lipid-binding regions, which impart the potential to interact with a variety of additional cytoskeletal regulating factors (Dominguez and Holmes, 2011; Haarer and Brown, 1990; Kaiser et al., 1999; Kovar et al., 2000; Krishnan and Moens, 2009; Thorn et al., 1997; Vinson et al., 1998).
There are four Profilin isoforms present in humans with ~ 33–95% similarity in amino acids (Fig. 2B–C). Depending on cell type, the estimate and measured amount of Profilin in mammalian cells is between 50 and 200 μM, which is generally considered sufficient or even excessive to the concentration of actin monomers in non-muscle cells (Funk et al., 2019; Kaiser et al., 1999; Lämmermann and Sixt, 2009; Vargas et al., 2017). Profilin-1 and Profilin-2 are ubiquitously expressed and the most abundant Profilin homologs in mammalian cells, whereas Profilin-3 and Profilin-4 proteins and transcripts have only been detected in the testes and kidneys (Behnen et al., 2009; Fagerberg et al., 2014; Mouneimne et al., 2012; Witke, 2004; Witke et al., 1998). Profilin-1 is the best characterized isoform, which is a ligand for over 128 binding partners and plays roles in actin assembly, nuclear transport and DNA repair, phase-separation, and as a tumor suppressor in diverse cancers (Burke et al., 2014; Henty-Ridilla and Goode, 2015; Hurst and Welch, 2011; Percipalle and Vartiainen, 2019; Posey et al., 2018; Rotty et al., 2015; Stüven et al., 2003; Suarez et al., 2015; Suarez and Kovar, 2016; Wittenmayer et al., 2004).
Finally, while much focus (> 2000 research articles) has illuminated the role of Profilin isoforms with regard to cellular processes that require proper actin dynamics for function, evidence suggests Profilin isoforms can regulate properties of microtubule dynamics through direct and indirect mechanisms in cells and reconstituted assays (see below) (Henty-Ridilla et al., 2017; Holzinger et al., 2000; Nejedla et al., 2016; Witke et al., 1998). Both Profilin-1 and Profilin-2 have been localized to microtubule structures in cells. Profilin-2 has been found on mitotic spindles and asters during mitosis, and Profilin-1 remains present on the sides of microtubules in mouse melanoma cells following cytoplasmic extraction (Di Nardo et al., 2000; Nejedla et al., 2016). Together these findings suggest an intriguing and emerging role for Profilin isoforms in regulating actin-microtubule crosstalk mechanisms (Henty-Ridilla et al., 2017; Holzinger et al., 2000; Pinto-Costa and Sousa, 2019; Wu et al., 2012). Such interactions and isoform differences may ultimately dictate a protective or malignant function in many pathologies including allergies, cardiovascular disease, Amyotrophic Lateral Sclerosis (ALS), Huntington’s Disease, Alzheimer’s Disease, diabetes, and lung, breast, and colorectal cancers (Ali et al., 2016; Alvarado et al., 2014; Bae et al., 2009; Boopathy et al., 2015; Burnett et al., 2008; Caglayan et al., 2010; Chakraborty et al., 2014; Coumans et al., 2014; Hauser et al., 2010; Henty-Ridilla et al., 2017; Horrevoets, 2007; Janke et al., 2000; Kim et al., 2015; Kooij et al., 2016; Lee et al., 2005; Roy and Jacobson, 2004; Santos and Van Ree, 2011; Shao et al., 2008; Tang et al., 2015; Wittenmayer et al., 2004; Wu et al., 2012; Zoidakis et al., 2012).
3. BIOCHEMISTRY OF PROFILIN IN ACTIN AND MICROTUBULE DYNAMICS
Competitive interactions underpin all biological processes ranging from individual molecules to ecological scales. Profilin is a champion of navigating the biochemical competition to ultimately regulate actin and microtubule dynamics, lipid and receptor interactions at the membrane, and diverse aspects of cancer signaling. Attempts to engineer a completely functional tagged version of Profilin to assess interactions between > 100 cellular binding partners or actin monomers have been extremely challenging (Funk et al., 2019; Melak et al., 2017; Plastino and Blanchoin, 2018; Skruber et al., 2018, 2020; Vitriol et al., 2015). Conventional fluorescent tags (~ 30 kDa) are twice as large as Profilin (~ 15 kDa). Tagging Profilin on the C-terminal end disrupts PLP-binding and using the N-terminus or an exposed loop can reduce Profilin-actin binding interactions (Nejedla et al., 2017; Wittenmayer et al., 2000). Direct labeling approaches for detecting single molecules of Profilin require four to five mutations and can result in aggregation (Henty-Ridilla et al., 2017; Vinson et al., 1998). As a consequence, most of what we know about Profilin has been determined from biochemical investigations of purified proteins, rather than cellular observations.
Cellular actin exists in several forms that are generated and maintained by different classes of regulatory proteins: straight filaments, bundles, branched networks, and globular monomers (Blanchoin et al., 2014; Michelot and Drubin, 2011; Svitkina, 2018; Svitkina and Borisy, 1999). Based solely on biochemical principles (the critical concentration for assembly is 0.1 μM), and estimates of the total cellular concentration of actin (~ 20–200 μM) and Profilin (20–200 μM), actin in cells predominately exists as polymerized filaments (Dominguez and Holmes, 2011; Funk et al., 2019; Gordon et al., 1977; Lodish et al., 2000; Pollard, 1986; Pollard et al., 2000). Actin filaments are constantly remodeled, assembled, destroyed, and recycled to reflect changing cell requirements in response diverse stimuli (Henty-Ridilla et al., 2013; Pollard, 2016). To achieve such actin dynamics, cells maintain a substantial stash of monomeric actin (~ 50–90% of the total actin depending on cell source), buffered by sequestering proteins that include Profilin and Thymosin-β4 (Tβ4) (Carlsson et al., 1977; Dominguez and Holmes, 2011; Pollard et al., 2000; Safer et al., 1991; Skruber et al., 2018). It has been tempting to speculate that most monomeric actin in cells exists in a 1:1 complex with Profilin because the concentration is commonly reported as near equimolar with actin, Profilin effectively out-competes Tβ4 for binding actin monomers, and structurally inhibits new actin filament polymerization (Blanchoin et al., 2014; Carlier et al., 1996; Carlsson et al., 1976b; Cooper et al., 1984; Dominguez and Holmes, 2011; Funk et al., 2019; Goldschmidt-Clermont et al., 1992; Kaiser et al., 1999; Vinson et al., 1998). Although the activities of Profilin sum to an attractive explanation for the requirement of nucleation factors to stimulate actin assembly, the balance between free actin monomers and those bound to Profilin, Tβ4, or other cellular ligands is not well characterized and extremely challenging to determine (Aroush et al., 2017; Henty-Ridilla et al., 2017; Plastino and Blanchoin, 2018; Skruber et al., 2018). Further, the availability of cellular actin monomers also fluctuates due to mechanisms of actin filament turnover and the presence of suites of disassembly proteins that synergize with Profilin to convert actin filaments into a polymerizable or sequestered monomeric state (Aroush et al., 2017; Blanchoin et al., 2000; Blanchoin and Pollard, 1998; Brieher, 2013; Bubb et al., 2003; Didry et al., 1998; Pollard, 2016; Pollard and Borisy, 2003; Pollard et al., 2000; Yarmola et al., 2001).
Profilin has a profound effect on actin polymerization in other ways besides sequestering actin including recharging actin monomers with ATP for new filament assembly, interactions with signaling lipids and proline-rich proteins (Chou and Pollard, 2019; Machesky and Pollard, 1993; Merino et al., 2018; Schlüter et al., 1997). Only two proteins are known to catalyze nucleotide exchange on actin monomers, Cyclase-Associated Protein (CAP) and Profilin. Historically Profilin has been depicted as the key driver of actin monomer recycling because it is able to bind the “aged” ADP-bound actin monomers that are released from depolymerizing actin filaments (Bertling et al., 2007; Goldschmidt-Clermont et al., 1992; Kotila et al., 2018; Mockrin and Korn, 1980; Vinson et al., 1998; Wolven et al., 2000). New evidence suggests that CAP can do this more efficiently than Profilin in mammalian systems, and there are organisms where Profilin is not known to catalyze this nucleotide exchange (Kotila et al., 2018; Ono, 2013). Profilin also has strikingly different affinities for monomers and for the growing ends of polymerizing actin filaments, and this property permits the efficient dissociation of Profilin from the ends of growing filaments (Courtemanche and Pollard, 2013; Funk et al., 2019; Jégou et al., 2011). Profilins have high affinity for binding PIP lipids which are conveniently located in the cellular membranes that polymerizing actin filaments push against to generate forces, create protrusions, and drive endocytosis (Dürre et al., 2018; Lu and Pollard, 2001). Interactions between PIP lipids and Profilin facilitate the release of Profilin-bound actin monomers for actin assembly (Lassing and Lindberg, 1985; Lu and Pollard, 2001; Ostrander et al., 1999; Schlett, 2017). All Profilins are known to bind some form of PIP lipid alone or in complex with actin monomers, with the exact PIP binding site defined by ~ 11 positively charged amino acids (Lu and Pollard, 2001; Sohn et al., 1995; Vinson et al., 1993). It is difficult to compare the exact affinities of Profilin for various PIP lipids across studies because different experimental methods (i.e., PIP compositions, ionic strength buffers, and wash steps) have been used and each of these variables drastically influences PIP binding (Shirey et al., 2017).
Profilin can reliably recognize and jointly bind, actin monomers and proline-rich motifs (i.e., PLP) present in many cytoskeletal regulatory proteins including: Formins, Ena/VASP, WASP/VCA-domain activators of the Arp2/3 Complex, and Drebrin (Chang et al., 1997; Evangelista, 1997; Evangelista et al., 2002; Ferron et al., 2007; Gertler et al., 1996; Higgs and Pollard, 1999; Mammoto et al., 1998; Miki et al., 1998; Reinhard et al., 1995; Rodal et al., 2003; Suetsugu et al., 1998). The ability to simultaneously bind actin monomers and PLP motifs endows Profilin with strong regulatory powers over actin polymerization and the actin monomer pool. Profilin-Formin interactions are mediated by PLP tracks present in all Formin proteins (Paul et al., 2008). Profilin interactions with specific Formin PLP tracks can strongly enhance the elongation and nucleation phases of actin assembly with micromolar to millimolar affinities (Chang et al., 1997; Courtemanche, 2018; Funk et al., 2019; Horan et al., 2018; Paul and Pollard, 2008; Perelroizen et al., 1994; Petrella et al., 1996; Sagot et al., 2002; Sherer et al., 2018; Watanabe et al., 1997; Zweifel and Courtemanche, 2020). The number and lengths of PLP tracks vary with each Formin, but each track competes to bind Profilin-bound actin monomers which ultimately increases the probability and speed actin monomers will be added to the growing actin filament in the correct orientation (Courtemanche and Pollard, 2012; Horan et al., 2018; Sherer et al., 2018; Zweifel and Courtemanche, 2020). Profilin-bound actin monomers hinder branched actin filament assembly mediated by the Arp2/3 Complex which prefers unbound actin monomers for assembly (Burke et al., 2014; Mullins et al., 1998; Rodal et al., 2003; Rotty et al., 2015; Skruber et al., 2018, 2020; Suarez et al., 2015; Suarez and, Kovar, 2016). Contrary to this mechanism, actin assembly mediated by Formin proteins requires Profilin and is significantly enhanced with Profilin-bound actin monomers (Burke et al., 2014; Chang et al., 1997; Evangelista et al., 2002; Funk et al., 2019; Henty-Ridilla and Goode, 2015; Kovar et al., 2003; Neidt et al., 2009; Romero et al., 2004; Rotty et al., 2015; Skruber et al., 2018, 2020; Suarez et al., 2015; Suarez and Kovar, 2016). Thus, competition between different actin nucleation systems has led the popular idea that Profilin tunes specific forms of actin assembly depending on the concentration of active nucleation proteins present (Rotty et al., 2015; Skruber et al., 2020; Suarez et al., 2015). Profilin-Formin isoform pairs in worms can further tune these activities (Neidt et al., 2009), which may have important implications in systems with higher numbers of Formin and Profilin isoforms present.
While much attention has focused on the role of Profilin in regulating actin dynamics, Profilin is also capable of regulating microtubule polymers and actin-microtubule crosstalk. In one of the first comprehensive studies comparing Profilin isoforms, tubulin and microtubule-associated proteins were first identified as ligands of Profilin-1 and Profilin-2 from affinity chromatography of mouse brain extracts (Witke et al., 1998). Profilin directly binds to microtubule sides (KD = ~ 11 μM) through specific amino acids in sites adjacent to the actin-binding surface on Profilin, and this microtubule binding activity is sensitive to the presence of actin monomers when both cytoskeletal elements are present in equal concentrations (Henty-Ridilla et al., 2017). In cells, Profilin resides on spindle and astral microtubules during mitosis and influences microtubule dynamics (Di Nardo et al., 2000; Henty-Ridilla et al., 2017; Nejedla et al., 2016). Some microtubule effects may be indirectly mediated through interactions between Profilin and Formin proteins that can also bind to microtubules (Bender et al., 2014; Nejedla et al., 2016; Pinto-Costa and Sousa, 2019; Szikora et al., 2017). At present there is not a simple assay to assess whether endogenous Profilin influences microtubule dynamics through direct mechanisms in cells. However, based on biochemical observations, cellular concentrations, estimates of the size of the Profilin-bound actin monomer pool, and relevant protein affinities, it is very likely that a pool of free “unbound” Profilin exists in the cytoplasm of mammalian cells and is available to bind microtubules and additional ligands at physiological concentrations (Fig. 3) (Henty-Ridilla et al., 2017; Henty-Ridilla and Goode, 2015; Plastino and Blanchoin, 2018).
Fig. 3.
Competition for Profilin Between Cellular Ligands Dictate the Types of Cellular Cytoskeletal Structures Formed. Cartoon model for the distribution of Profilin to actin, microtubules, or regulatory ligands (Formins, Ena/VASP, the Arp2/3 Complex). Based on biochemical principles, free Profilin pools likely exist in cells. Direct interactions between isoforms of Profilin and tubulin are hypothesized but not yet directly confirmed (Henty-Ridilla et al., 2017; Nejedla et al., 2016; Pinto-Costa and Sousa, 2019; Witke et al., 1998).
4. ROLE OF PROFILIN ISOFORMS IN CANCER
Humans have four Profilin isoforms, with Profilin-1 commonly accepted as is the most ubiquitous and abundant isoform in almost all tissues and cell types (Fig. 4A) (Behnen et al., 2009; Fagerberg et al., 2014; Mouneimne et al., 2012; Witke, 2004; Witke et al., 1998). Thus, the majority of cellular and biochemical studies have focused on the activities of Profilin-1. Profilin-3 transcripts are virtually absent from all tissues except kidneys where transcripts are 83-fold less abundant than Profilin-1 (Fig. 4A). Profilin-4 transcripts are more abundant than Profilin-3 across tissues except kidneys, but are still much less abundant than either Profilin-1 or Profilin-2 isoforms (Fig. 4A). The only known location where Profilin-1 is not the most predominate isoform is in neuronal-derived tissues and cells. Here, Profilin-2 proteins and transcripts have been measured ~ 5-fold more abundant than Profilin-1, although the exact mechanisms that underlie this distinct distribution are still not fully elucidated (Fig. 4A) (Gareus et al., 2006; Mouneimne et al., 2012; Witke et al., 1998). There are two alternatively spliced versions of Profilin-2 (designated 2a and 2b) differing by nine amino acids in the C-terminal region and an extended patch of aromatic resides (Gieselmann et al., 2008; Lambrechts et al., 1997; Nodelman et al., 1999)Both. splice variants of Profilin-2 have similar affinities for actin but differ in binding other ligands (Nodelman et al., 1999; Witke et al., 1998). Profilin-2a is the predominant form, whereas Profilin-2b is restricted to very limited tissues (Lambrechts et al., 2006). While Profilin-1 and Profilin-2 have similar effects on many biochemical properties pertaining to actin dynamics including nucleotide exchange and binding phosphatidylinositol (PIP) lipids, Profilin-2 has a five-fold lower binding affinity for actin monomers and has higher affinity for EVL and VASP PLP-containing ligands (Gieselmann et al., 2008; Mouneimne et al., 2012). Thus far, Profilin-2 has been studied as a regulator of actin dynamics in PLP ligand binding assays.
Fig. 4.
Profilin Isoform Transcripts in Normal and Tumor Tissues. (A) Transcript levels for Profilin-2, Profilin-3, and Profilin-4 relative to the most ubiquitous expressed isoform, Profilin-1. We normalized the means of all available RNAseq transcript data in “Projects” that contained normal tissue samples from cancer patients (n = 1–215) currently available in the National Cancer Institute Genomic Data Commons (https://portal.gdc.cancer.gov; Grossman et al., 2016). The purple dotted line represents Profilin-1 levels in each tissue shown. AG, adrenal gland. HNSC, head neck squamous cells. (B) Mean RNA transcripts of Profilin-1 and Profilin-2 obtained from the database in (A) for normal and tumor tissues in fragments per kilobase exon model per million mapped reads (FPKM), (n = 1–1191). Error bars, SEM. (C) A Venn diagram summarizing the changes found in (B) for Profilin-1 (PFN1) and Profilin-2 (PFN2) transcripts in tumors compared to normal tissues. (D) Fold change in the Profilin-1:2 ratio in primary tumors. All data was downloaded and analyzed from https://portal.gdc.cancer.gov on 10 April 2020 (data release 23.0). All transcript data was sorted by tissue source across “Project” databases with the exception of HNSC which was is presented a mix of tissue sources.
Several studies have suggested the intriguing and controversial idea that perturbations to Profilin-1 and Profilin-2 have opposing phenotypes in several cancers (Baraniskin et al., 2012; Cui et al., 2016; Janke et al., 2000; Mouneimne et al., 2012; Wittenmayer et al., 2004; Zhang et al., 2018; Zoidakis et al., 2012). In short, elevated levels of Profilin-1 are correlated with a tumor suppressive effect and reduced metastasis in breast, lung, colorectal, bladder, esophageal, and thyroid cancers, whereas elevated levels of Profilin-2 produce higher metastatic behaviors (Janke et al., 2000; Jiang et al., 2017; Mouneimne et al., 2012; Wittenmayer et al., 2004; Zhang et al., 2018; Zou et al., 2007). There are inconsistencies to this dichotomy, however, and no satisfying explanation reconciling the differences between these studies has been reached beyond differences arising from tissue-derived or cell line specific phenotypes and sample sizes used.
To attempt to clarify whether Profilin isoform levels are correlated with tumor development and/or metastasis across diverse cancers, we analyzed all RNAseq transcript data in “Projects” that contained normal tissue and primary tumor samples from cancer patients currently available in the National Cancer Institute Genomic Data Commons (https://portal.gdc.cancer.gov; Grossman et al., 2016) (Fig. 4). In most of the tissues examined (i.e., bladder, breast, esophagus, head and neck squamous cell (HNSC), lung, prostate, thymus, and uterus), both Profilin-1 and Profiln-2 transcript levels were elevated in primary tumors compared to normal tissues (Fig. 4B and C). The second most common trend observed was that primary tumor transcripts displayed decreased Profilin-1 transcripts and elevated Profilin-2 transcripts compared to normal tissues (i.e., adrenal gland, cervix, colon, pancreas, and skin) (Fig. 4B and C). The only measurement where both Profilin isoforms decreased was in stomach primary tumors (Fig. 4B and C). We did not observe any transcript-based trends between Profilin isoforms for metastatic and reoccurring tumors, although this data was much less abundant and not available for all the tissues investigated above. The most spectacular observation was in brain tumors—in normal tissue Profilin-2 transcripts are ~ 5-fold more abundant than Profilin-1; however, in primary tumors Profilin-1 transcripts outnumber Profilin-2 by 2.7-fold (Fig. 4B and C). Combined this results in an over 12-fold relative increase Profilin-1 transcripts in brain tumors!
These observations and previous studies may raise the exciting possibility that cellular ratios of Profilin-1 and Profilin-2 underlie the behavioral differences between cancers (Fig. 4D). In general terms the ratio of Profilin-1 to Profilin-2 transcripts in primary tumors decreases (Fig. 4D), but this is not simply the result of a change in Profilin-1 transcripts (i.e., Profilin-1 transcripts can be reduced, Profilin-2 transcripts can become elevated, or combination of both may occur) (Fig. 4B and C). We also explored whether patient survival was connected to changes in Profilin transcripts or ratios from this data, but no specific correlation was observed. However, this observation has several confounding factors that should be considered including: the similarities and differences in patient treatment plans, age, gender, and progression at diagnosis. Finally, while transcript levels are relatively easy to obtain or measure, they do not always reflect the amount of protein present and available for cellular activities, particularly those associated with regulating the cytoskeleton. Future studies will likely have to measure these parameters in specific cell types and circumstances.
5. PROFILIN IN CELL DIVISION
Cancer in the most basic sense is disordered or uncontrolled cell division instigating changes in the rate cells divide, the activity of cell cycle regulators and signals, or inhibition of normal cell maintenance/death. In a cancer-free context, the details underling the dynamics of the mitotic spindle have been the subject of intense scrutiny for literally hundreds of years (McIntosh and Hays, 2016). Changes in cell architecture supported by the actin and microtubule cytoskeletons is a normal requirement to progress through the cell cycle. Some of the most dramatic cytoskeletal reorganizations are triggered by Cyclin complexes which alter the dynamics of motor proteins organizing the mitotic spindle, Rho GTPases, and proteins that facilitate actin-microtubule interactions (Blangy et al., 1995; Böttcher et al., 2009; Jiang et al., 2015; Kita et al., 2019; Miller, 2011; Plessner et al., 2019; Pollard and Wu, 2010; Ubersax et al., 2003; Yamashiro et al., 1991). Much information that clarifies the roles for microtubules and many signaling factors is known, however functions of actin (and regulatory partners) in cell division have been more elusive and are just starting to emerge. For example, many tools used for visualizing cytoskeletal proteins in mitotic spindles were considered technically limiting and challenging for actin filaments, despite plentiful descriptions of its presence associated with the mitotic spindle (Cande et al., 1977; Gawadi, 1971; Herman and Pollard, 1979; Sanger, 1975). While the classic and most characterized roles for actin in cell division pertain to generating the forces required for cytokinesis, additional studies demonstrate that actin is important for positioning the spindle and spindle pole separation (Miller, 2011; Pelham and Chang, 2002; Rosenblatt et al., 2004; Théry et al., 2005; Toyoshima and Nishida, 2007; Watanabe et al., 2008), extensively reviewed: (Pollard and O’Shaughnessy, 2019). Recent observations show that there are dynamic populations of actin and actin-microtubule-associated structures localized to the mitotic spindle, and these structures reorganize as cells advance through the mitotic phase (Kita et al., 2019; Plessner et al., 2019). Actin nucleation proteins, microtubule-associated proteins, and Profilin are proposed mediators of these dynamics, but the detailed mechanisms of how they may go wrong in cancer are not clear (Henty-Ridilla et al., 2016, 2017; Ishizaki et al., 2001; Kita et al., 2019; Nejedla et al., 2016; Plessner et al., 2019; Roth-Johnson et al., 2014; Wade, 2007).
The role for Profilin in the cell cycle has been difficult to discern, complicated by its affinity for a plethora of signaling molecules and actin and microtubule regulation proteins. Classic genetics has demonstrated that Profilin-1 is essential for generating the cytokinetic ring, cell survival, and division in many organisms (Chang et al., 1997; Kandasamy et al., 2002; Kovar et al., 2003; Polet et al., 2006; Severson et al., 2002; Vidali et al., 2007; Witke et al., 2001). Further, in systems where multiple isoforms of Profilin are available, the loss of the most ubiquitous Profilin cannot be fully rescued for cell cycle effects with the other isoforms (Polet al., et al., 2006; Witke et al., 2001). Specialized mouse cells lacking Profilin-1 displayed morphological defects and aberrant actin filament distributions but were able to complete mitosis albeit on a slower timescale than normal cells (Böttcher et al., 2009). These phenotypes were unable to be rescued with Profilin point mutants deficient for binding actin or Formin proteins. These results may indicate cell dependent differences in Profilin function but may suggest the involvement of actin-independent functions of Profilin.
In addition to organizing microtubules, centrosomes organize a network of actin filaments generated by the Arp2/3 complex. Intriguingly, increasing density or crosslinking of actin filaments correlates with a reduction in microtubules in vitro and at centrosomes in cells (Colin et al., 2018; Farina et al., 2016, 2019; Inoue et al., 2019; Ricketts et al., 2019). Hence, the centrosome is a coordinator of actin and microtubules, and in light of biochemical studies elucidating the role of Profilin with actin nucleation promoting factors and microtubules, this relationship may be indirectly regulated through Profilin (Burke et al., 2014; Funk et al., 2019; Rotty et al., 2015; Skruber et al., 2020; Suarez et al., 2015). In sum, the complete functions of Profilin in cell division remain unclear. Some roles are likely directly related to Arp2/3- or Formin-mediated actin filament assembly or in regulating microtubule dynamics, and some may go beyond including generating the forces for cytokinesis (Chang et al., 1997; Kita et al., 2019; Kovar et al., 2003; Nejedla et al., 2016; Plessner et al., 2019; Severson et al., 2002).
6. PROFILIN IN CELL MOTILITY AND METASTASIS
The migration of cells is a complex biological process that requires the reorganization of actin, microtubules, membrane receptors, lipids, and the cell-matrix. The loss of proper cell migration has profound effects on neuronal pathfinding, development, wound healing, and overactive migration is a classic hallmark of metastasis and ultimately responsible for distributing tumorigenic cells to sites in the body where they do not normally belong. Cells initiate movements by extending lamellar membrane protrusions driven by physical forces produced by assembling actin filaments that push on the membrane surface. Traction is produced from directionally elongated focal adhesion sites and actin stress fibers to propel the cell forward. Meanwhile contractile forces produced by actin-myosin stress fibers retract the back of the cell as it advances onward (Pollard and Borisy, 2003). Less is known about microtubules or crosstalk between actin and microtubules in this process, however pharmacological evidence demonstrates that perturbing either system alters motile behaviors (Coles and Bradke, 2015; Dogterom and Koenderink, 2019; Etienne-Manneville, 2004; Rodriguez et al., 2003). There are also numerous connections between microtubules and focal adhesion complexes as well as integrin-based extracellular matrix adhesions (Borisy et al., 2016; Bouchet and Akhmanova, 2017; Bouchet et al., 2016; Dziezanowski et al., 1980; Euteneuer and Schliwa, 1984; Kaverina and Straube, 2011; Kaverina et al., 1998; Rodionov et al., 1998; Wittenmayer et al., 2004; Wittmann and Waterman-Storer, 2001).
Motility can be recapitulated on a bead in vitro or in genetically tractable motile organisms like Listeria from a core set of proteins including actin, the Arp2/3 Complex, Capping Protein, Cofilin, and Profilin (Loisel et al., 1999; Pantaloni et al., 2001; Pollard and Borisy, 2003; Theriot et al., 1992, 1994; Tilney et al., 1992; Wiesner et al., 2003). Actin filaments at the leading edge are mostly formed by the Arp2/3 Complex oriented with the faster growing end oriented outward (Pollard and Borisy, 2003; Rouiller et al., 2008; Small, 1988; Small and Celis, 1978; Svitkina, 2018; Svitkina and Borisy, 1999; Svitkina et al., 2003; Symons and Mitchison, 1991). Growing filaments are capped relatively quickly at short lengths and are therefore mechanically suited to generate/sustain sufficient forces to propel the cell forward (Akin et al., 2008; Blanchoin et al., 2000; Mogilner and Oster, 1996). Biochemical evidence demonstrates that the presence of Profilin-bound actin inhibits branched actin assembly, that Profilin is required for nucleotide exchange to assemble new actin filaments, and that amounts of free Profilin can compete for the faster growing end of actin filaments with Capping Proteins, Formins, and other ligands (Bubb et al., 2003; Cooper et al., 1984; Dos Remedios et al., 2003; Kaiser et al., 1999; Mockrin and Korn, 1980; Mullins et al., 1998; Pernier et al., 2016; Rotty et al., 2015; Skruber et al., 2020; Suarez et al., 2015; Vinson et al., 1998). There are still many questions that underlie the behavior of actin and microtubule networks at the leading edge of crawling cells. How do cells assemble and rearrange cytoskeletal polymers so quickly? Why do cells expend so much energy incessantly constructing and disassembling these proteins? What detailed roles do the five minimal proteins perform in cells, can they be visualized, and how do they go awry in disease?
7. CELL SIGNALS CONVERGING ON PROFILIN
Many interconnected signaling pathways and feedback loops contribute to cell homeostasis and respond in disease, particularly in cancer. This labyrinth of signals commonly challenges the development of therapeutics, especially when target molecules exhibit high sensitivity to a diversity of ligands spanning pathways on very rapid timescales. Several signaling pathways converge on Profilin in cancer (Fig. 5), however whether or not these pathways use Profilin in its roles as a regulator of the cytoskeleton are not always clear. In addition, specific modifications to Profilin (usually phosphorylation) directly impact actin dynamics, but the identity and timing of signals and kinases regulating these modifications in disease remain unknown.
Fig. 5.
Signaling Pathways Converge on Profilin. (A) Upon stimulation by an activating ligand the TGFβ signaling pathway activates Smads which stimulate the epithelial to mesenchymal transition (EMT). Profilin-2 has an inhibitory effect on this pathway by preventing HDAC1 signals to the nucleus. TF, transcription factor. (B) Upon activation by diverse receptor tyrosine kinases the P13K/TOR/AKT pathway ultimately stimulates cell proliferation, invasion, and migration. Phosphorylation of Profilin-1 (S137) loses the ability to bind actin monomers and becomes translocated to the nucleus and inhibits cell death pathways. (C) Activation of the VEGFR pathway leads to the phosphorylation of Profilin-1 (Y129), stimulating angiogenesis. Phosphorylated Profilin-1 (Y129) enhances nucleotide exchange on actin monomers, ultimately increasing overall actin filament polymerization. Numbers present throughout the figure are references (purple box) for particular signaling steps.
The Transforming Growth Factor Beta (TGFβ) pathway is essential in development, regulating cell growth and differentiation, apoptosis, and is a common place where signals go astray in diverse cancers, ultimately driving the epithelial to mesenchymal transition (EMT) of cancer cells and permitting invasive migratory behaviors (Moustakas and Heldin, 2008). TGFβ ligands bind cell receptors which recruit and phosphorylate signal transducing transcription factors (SMADs) to mediate downstream responses. While Profilin-1 has no reported effect on TGFβ signals, increased Profilin-2 protein correlates with SMAD2/3 signals, reducing Profilin-2 or SMAD2/3 levels correlated with tumor suppression in mice, and an early spike in TGFβ activity in a luciferase assay was reduced (Tang et al., 2015). Additional analysis revealed a downstream cytoplasmic interaction between Profilin-2 with HDAC1 that further reinforces SMAD nuclear activities by inhibiting HDAC1 (Fig. 5A) (Tang et al., 2015). The P13K/AKT/mTOR (Phosphatidylinositol 3-kinase/Protein Kinase B/mammalian Target of Rapamycin) intracellular pathway is one of the most commonly mutated in cancer (Melamed et al., 2019; Paplomata and O’Regan, 2014). The overexpression of Profilin-2 in head and neck cancer cell lines increased cell proliferation, and these cells displayed higher levels of phosphorylation for AKT and downstream effectors including β-catenin (Zhou et al., 2019). Therefore, Profilins are an important link in the web of cancer signaling pathways that require the cytoskeleton for function.
Protein phosphorylation is widely used to regulate biological functions. Phosphorylation is the only known category of post-translational modification of Profilin, and it can occur at several amino acid sites to regulate actin-based activities (Alkam et al., 2017). In cells, Profilin-1 can be targeted for phosphorylation by Protein Kinase C (PKC) at S137 downstream of P13K signals, and also by Rho-associated Kinase-1 (ROCK1) downstream of GTP-signals, and is linked to promoting metastasis and invasion in breast cancer cells (Fig. 5B) (Hansson et al., 1988; Rizwani et al., 2014; Sathish et al., 2004; Shao et al., 2008; Singh et al., 1996; Yang et al., 2017; Yao et al., 2014). In biochemical assays phosphomimetic mutations at this site indicate that actin monomers do not bind this modification of Profilin as efficiently compared to wild-type versions (Shao et al., 2008). Profilin-1 can also be phosphorylated at Y129, which is present in the actin binding site of the protein and unsurprisingly reduces the binding capacity of Profilin-1 for monomeric actin (Fan et al., 2012). In cells this modification occurs by Src kinase initiated through a Vascular Endothelial Growth Factor Receptor Kinase 2 (VEGFR2) signaling cascade (Fan et al., 2012; Simons and Schwartz, 2012). The Y129 phosphorylation has also been linked to the progression of glioblastoma by forming a complex with the Von Hippel-Lindau protein that prevents the degradation of hypoxia induced factor 1 alpha (HIF-1α), ultimately driving the vascularization of tumors and cancer progression (Fig. 5C) (Fan et al., 2014). The impacts of phosphorylation on Profilin-2 have not been as extensively investigated. Using an in silico approach, 14 potential phosphorylation sites on the Profilin-2a protein have been identified, seven of which were biochemically characterized: Y29, S71, S76, Y78, S129, Y133, and S137 (Walter et al., 2020). Phosphorylation of Profilin-2 at S71, S76, or S129 disrupted actin-binding activities, and intriguingly phosphorylation of S76 was able to stimulate the elongation phase of actin polymerization (Walter et al., 2020). Lastly, to use post-translational modifications as effective molecular switches, the cellular balance between the phosphorylated and dephosphorylated states of Profilin must be maintained. To date Protein Phosphatase-1 (PP1) is the only known kinase to effectively dephosphorylate Profilin-1, specific to amino acid S137 (Shao and Diamond, 2012). Whether PP1 can perform this role at other phosphorylation sites in Profilin-1 or Profilin-2 or if other kinases contribute to these functions has not been fully explored.
8. PROFILIN IN IMMUNE SYSTEM RESPONSES
How do cancer cells avoid detection or eradication by the immune system? Can targeting the host immune responses contribute to better treatment outcomes? The presence of inflammatory immune cells in human tumors and innate (receptor-ligand interactions) and adaptive immune responses (phagocytic macrophages) contribute to the progression of cancer by inducing immunosuppression, stimulating cancer proliferation and metastasis (Palucka and Coussens, 2016). Reports have investigated the contribution of Profilin in the innate and adaptive immune responses elicited by different microbes, but these responses in the context of cancer or cancer recovery have not been explored. In the human adaptive immune system, actin and microtubules are essential for migration, phagocytosis, cell secretion, and cell-cell interactions (Mostowy and Shenoy, 2015; Pfajfer et al., 2018; Wickramarachchi et al., 2010). Although only disruptions to actin dynamics were investigated, cells without Profilin-1 or Profilin-2 fail to perform actin-microtubule mediated phagocytosis in macrophages (Coppolino et al., 2002; Kim et al., 2012). In Cytotoxic T Lymphocytes (CTLs) Profilin negatively regulates the exocytosis of lytic granules, which may suggest an enhanced ability to both eliminate tumor cells and increase the migration and invasion of these “helpful” immune cells (Schoppmeyer et al., 2017). In other immune cells (e.g., dendritic cells, neutrophils) Profilin-1 protein levels were higher than cancer cell lines (HT1080 and B16F10) and Profilin-2 was only detected in the dendritic cell line used (Funk et al., 2019). Presumably these findings suggest a role for Profilin in ameboid migratory behaviors controlled by the actin and microtubule cytoskeletons but the exact mechanisms have not been fully elucidated (Lämmermann and Sixt, 2009).
9. TARGETING PROFILIN AS A CANCER THERAPEUTIC
The timing of the cell cycle, the morphology of cells and tissues, and the direction and speed of cellular movements are essential processes regulated by the broad actions of cellular actin and microtubule dynamics. As a consequence, compounds (synthetic and natural) that disrupt cytoskeletal dynamics are among the most widely utilized chemotherapeutics available. These properties also cause treatments to be extremely toxic to patients. With the critical nature of microtubules in cells, drugs that target microtubule dynamics are some of the most effective therapeutics available (Mukhtar et al., 2014). For example, one of the first compounds that targeted the cytoskeleton to treat cancer was Taxol, which stabilizes microtubules and effectively arrests cell division in breast, ovarian, lung, prostate, blood, and many other cancers (Fife et al., 2014; Weaver, 2014). Existing pharmaceutical agents target actin dynamics (i.e., latrunculins, cytochalasins, jasplakinolides), however these compounds are indiscriminately toxic to numerous organs, cardiac, and muscle function in addition to cancerous tumors (Bonello et al., 2009). The development of compounds that target actin and microtubule regulatory proteins are even more rare, but come with the advantage of a targeting specific features of cytoskeleton dynamics (i.e., cytoskeletal assembly, disassembly, stabilization, turnover, or motor protein dynamics) or potentially actin-microtubule crosstalk. Compounds targeting the microtubule-associated proteins Tau or the kinesin Eg5 are excellent at inducing mitotic arrest and limiting tumor proliferation but frequently fail as a clinical monotherapy due to their acute specificity (Chan et al., 2010; Engelke et al., 2016; Hancock, 2014; Milic et al., 2018; Pan et al., 2017; Smith et al., 2013; Sturgill et al., 2016). Other small-molecule screens targeting actin assembly identified inhibitors for the Arp2/3 complex, N-WASP, Tropomyosins, and Formins, although recent evidence questions the specificity of some of these molecules in cells (Bolger-Munro et al., 2019; Hetrick et al., 2013; Isogai et al., 2015; Nolen et al., 2009; Peterson et al., 2004; Rizvi et al., 2009; Sellers et al., 2020; Stehn et al., 2013).
Profilin may represent an effective therapeutic target to fight diverse cancers due to its roles in cytoskeletal regulation and position in cancer signaling cascades. Intriguingly hyperactive Profilin-1 in signaling pathways can lead to precocious apoptosis and resistance to several chemotherapeutics, while silencing Profilin-1 can reduce tumor growth in vivo (Frantzi et al., 2016; Zou et al., 2010). Small molecule screens have revealed compounds that mitigate breast cancer-induced changes in Profilin expression and migration and two small molecules that prevent the interaction of Profilin with actin monomers have been identified (C1 and C2) (Gau et al., 2018; Joy et al., 2014). In biochemical assays C1 and C2 obstruct Profilin-actin binding at high concentrations (50–100 μM), supporting more total actin polymerization than control assays conducted in their absence (Gau et al., 2018). In cells these compounds slowed endothelial cell migration, proliferation, and inhibited angiogenesis (Gau et al., 2018). Do these small molecules also target the actin binding affinity of other Profilin isoforms and actin? Further biochemical and cellular characterization of C1 and C2 with regard to cancer signaling, phosphorylation state, lipid-binding, or microtubule effects may provide valuable mechanistic insights for using these molecules to treat diseases.
10. TALES FROM DIVERSE MODEL SYSTEMS
Many mechanisms underlying protein function are conserved across evolutionary scales. The natural course of disease can sometimes take a lifetime to manifest (evolutionary lifetime is shorter in some organisms). Model organisms can quickly develop a disease or its symptoms allowing researchers to study links between genetic factors, aberrant protein functions and cellular processes on a much shorter time frame. Studying the diversity of protein homologs could lead us down unexpected paths illuminating new therapeutics or elucidating new molecular connections that can be exploited with new treatment approaches. Profilins are evolutionarily conserved in all forms of life where it has been investigated (including Archaea, bacteria, viruses, and eukaryotes) and this provides an exceptional opportunity to take advantage of model organisms to study its role as a regulator of actin and microtubule dynamics in disease.
There is much to learn about cancer not only by observing the differences between normal biology and how normal biology goes wrong, but also from how organisms use the same or similar biology in unique ways. Developmental regimes in the Drosophila, zebrafish, and C. elegans model systems are similar to cancer progression and metastasis in that they require the same tools for execution: cytoskeletal dynamics, cell migration, and cell division. The simplest mechanism explaining the connection of the development of each these model organisms with cancer is that that the loss or misregulation of Profilin is linked to failures in actin assembly and microtubule dynamics. Intriguingly these dynamics are not restored by other Profilin isoforms in vertebrates and many eukaryotes (Cooley et al., 1992; Kovar et al., 2000; Lai et al., 2008; Müssar et al., 2015; Polet et al., 2006; Reeve et al., 2005; Verheyen and Cooley, 1994; Witke et al., 2001). Profilin is critical for maintaining the mesh of actin bundles keeping Drosophila nurse cells (in oocytes) intact (Ghiglione et al., 2018; Verheyen and Cooley, 1994). In addition to reduced viability, flies with reduced Profilin levels possess a weakened microtubule spindle apparatus, a less contractile actomyosin cytokinetic ring, and over-proliferative somatic cells (Giansanti et al., 1998; Giansanti and Fuller, 2012; Shields et al., 2014; Verheyen and Cooley, 1994). Zebrafish and C. elegans require Profilin for the growth of neuronal cells, neuronal maturation, myelination, and muscle development (Ehler, 2018; Kooij et al., 2016; Kwak et al., 2013; LeCorgne et al., 2018; Majesky, 2007; Meyers, 2018; Murk et al., 2009; Polet et al., 2006; Roth et al., 1999; Yuan et al., 2018). Profilin from C. elegans is required for anterior-posterior establishment, DNA positioning and abscission during mitosis, and to optimally regulate Formin-mediated actin polymerization through specific Formin-Profilin isoform pairs (Davies et al., 2018; Neidt et al., 2009; Panzica et al., 2017; Severson et al., 2002). Zebrafish undergo extensive cell migration phases during development that require key regulators of actin for normal execution. Profilin is required for the completion of gastrulation, endothelial cell proliferation, neural cord development, and establishing heart progenitor cell lineage (Ding et al., 2006; Lai et al., 2008; Yuan et al., 2018). Thus, Profilin-mediated development has historically revealed fresh perspectives for the underlying mechanisms that explain how Profilin goes rogue in cancer from these organisms.
Several viral Profilin homologs effectively bind mammalian actin monomers but with weaker affinity and actin nucleotide exchange (Blasco et al., 1991; Butler-Cole et al., 2007; Machesky et al., 1994; Moreau et al., 2017, 2020). Neither Vaccinia nor Ectromelia homologs bind PLP regions (Butler-Cole et al., 2007; Machesky et al., 1994). Intriguingly, Vaccinia Profilin binds phosphoinositide (PIP) lipids with higher affinity than human Profilin-1, and Profilin from Ectromelia does not bind PIP lipids but directly interacts with other actin regulatory proteins like tropomyosin for function (Butler-Cole et al., 2007; Machesky et al., 1994). Apicomplexa are extremely susceptible to actin-polymerizing and depolymerizing agents (Baum et al., 2006; Gordon and Sibley, 2005). Toxoplasma gondii Profilin binds and sequesters actin monomers, and loss of Toxoplasma Profilin prevents parasite replication and host invasion by disrupting host and parasite actin dynamics (Plattner et al., 2008; Skillman et al., 2012). Genetic, immunological, structural, and cell biological studies have further demonstrated Toxoplasma Profilin is important in pathogen-host interactions initiated through interleukins and toll-like receptors from both host and parasite (Denkers, 2010; Kucera et al., 2010; Plattner et al., 2008; Yarovinsky et al., 2005). Similar observations converging on Profilin and actin dynamics in innate immunity have been investigated in chytrid fungi and plants (Babik et al., 2014; Cao et al., 2016; Qiao et al., 2019; Sun et al., 2018). Thus, actin regulation by Profilin proteins has been important in the “evolutionary arms race” between hosts and diverse microbes.
The role of cilia in cancer signaling and cell cycle regulation with regard to the microtubule cytoskeleton has been studied extensively (Fabbri et al., 2020; Goetz and Anderson, 2010; Golemis et al., 2018; Higgins et al., 2019). Emerging evidence demonstrating the involvement of actin in ciliary formation and development from the tractable model system Chlamydomonas reinhardtii demonstrate important new forms of ciliary regulation that challenge long-held ideas suggesting actin or actin-microtubule crosstalk is necessary for normal ciliary assembly, motility, and signaling. Chlamydomonas is an excellent model system for studying the duality of the actin and microtubule cytoskeletons in disease. It features two easily accessible cilia that behave and are regulated by mechanisms virtually identical to mammalian forms and has yielded important insights into human diseases and developmental disorders including primary ciliary dyskinesia (PCD), polycystic kidney disease (PKD), situs inversus, and numerous ciliopathies (Harris, 2001; Pazour and Witman, 2009; Wase et al., 2019). Chlamydomonas actin structures depend on specific actin architectures and localizations to execute diverse cell processes (Christensen et al., 2019; Craig et al., 2019; Detmers, 1985; Detmers et al., 1983; Harper et al., 1992; Jack et al., 2019; Kovar et al., 2001; Onishi et al., 2016; Piperno and Luck, 1979; Wilson et al., 1997). This includes cilia which were historically studied with microtubules as the predominant cytoskeletal polymer in ciliary assembly (Avasthi et al., 2014; Jack et al., 2019; Kovar et al., 2001; Tai et al., 1999). Chlamydomonas Profilin is found throughout the organism including ciliary structures (Kovar et al., 2001). Chlamydomonas Profilin binds actin, can inhibit aspects of Arp2/3-mediated actin assembly and enhance actin assembly through Formins. However, this Profilin is unique—it does not recycle nucleotides on actin monomers, it very potently inhibits spontaneous filament nucleation, it caps the fast-growing ends of actin filaments 5- to 10-fold more efficiently than other homologs, and seems to protect a specific actin isoform (IDA5) from degradation (Christensen et al., 2019; Courtemanche and Pollard, 2013; Kovar et al., 2001; Onishi et al., 2016; Pernier et al., 2016). Competitive interactions between Profilin and diverse actin assembly factors may ultimately dictate timing and dimensions of assembled actin in these cells. In addition, a single Profilin regulates two very different actin isoforms (IDA5 and NAP1). The occupation of actin filament ends by Profilin may limit the role for Profilin on ciliary microtubules or liberate shared actin-microtubule regulators to orchestrate linked cytoskeletal behaviors. Thus, Chlamydomonas, is uniquely situated to elucidate foundation mechanisms concerning the role of Profilin in actin assembly and as a facilitator of actin-microtubule crosstalk.
Yeast model systems are genetic powerhouses that have been indispensable in developing a “parts list” and interactome for many complex pathways and cellular processes. Many foundational studies dissecting Profilin-mediated cytoskeletal dynamics come from yeast, including the discoveries that: Profilin facilitates nucleotide exchange with Srv2/CAP (Amberg et al., 1995; Lila and Drubin, 1997; Ono, 2013; Wolven et al., 2000); Profilin interacts with essential actin assembly factors during cell division (Chang et al., 1997); Profilin synergizes with Formin proteins to promote actin polymerization (Pruyne et al., 2002; Sagot et al., 2002); and that Profilin dictates actin structure (straight or branched filaments) by regulating the actin monomer pool (Suarez and Kovar, 2016). The ability to precisely engineer yeast coupled with a rapid life cycle are unparalleled for dissecting mechanisms in vivo. In one compelling example, the interaction and mechanism of how Profilin binds to PLP motifs was dissected by introducing 87-point mutations into yeast Profilin (Lu and Pollard, 2001). Similar approaches have been used to quickly assess the viability and role of disease-specific Profilin variants in human disease (Figley et al., 2014). Further, yeast-based technologies have been utilized in drug discovery screens, to produce anti-cancer drugs, and personalized cancer therapies (Ferreira et al., 2019).
Profilins and many of their biochemical activities are conserved across evolution—features that provide an exceptional opportunity to employ model organisms in studying the roles of Profilin as a regulator of actin and microtubule dynamics in disease. A major drawback of conventional mammalian systems in such endeavors is the inability to resolve the fine details of cytoskeletal dynamics in vivo by the standard microscopy techniques used in many screens (i.e., interactor, small-molecule therapeutics, localization) and even fewer in living organisms. Traditional single-molecule attempts to visualize individual actin filaments require injecting fluorescently tagged actin polymers or though FRAP/photoconversion methods that track fiducial marks on preformed actin filaments or bundles (Dovas et al., 2011; McGrath et al., 1998; Wang, 1984; Waterman-Storer and Salmon, 1998). Plant model systems (Arabidopsis and Physcomitrella) may be the only model systems where dynamics of individual actin filaments have been measured in cells due to the presence of sparse cytoskeletal arrays that afford high-resolution on fast scales (ms) (Augustine et al., 2011; Staiger et al., 2009). Other model systems are approaching this resolution for actin and microtubules through creative combinations of traditional approaches (i.e., microinjection, FRAP, photoactivation), super-resolution imaging modalities, genetically-encodable fluorescently-stable single-molecule tags/tools, and gene-editing technologies (Aumeier et al., 2016; Fritzsche et al., 2017; Funk et al., 2019; Huang et al., 2008; Rust et al., 2006; Skruber et al., 2020; Tas et al., 2017; Vignaud et al., 2020; Vitriol et al., 2015).
11. OPEN QUESTIONS AND FUTURE DIRECTIONS
Profilin is a much more complicated and elegant molecule than suggested by its defining role as a sequestering protein. Profilin was first identified in the 1970s and since then thousands of publications have described and defined its mechanisms regulating the actin cytoskeleton. Recent studies suggest that Profilin still has secrets to share regarding how tubulin, microtubules, and cytoskeletal proteins are regulated, how Profilin is uniquely positioned to choreograph the cytoskeleton during essential cell processes, and the snafus that cause diseases including cancer, neurodegeneration, cardiovascular decline, allergies, and many more. We have still not discerned complete molecular mechanisms that connect Profilin to these processes. When it comes to the role of Profilin the details are truly important.
Clearly actin regulation is an essential function of Profilin, but what other roles do Profilin proteins fulfill? Some versions of Profilin contain extended structural regions, and although these regions do not alter Profilin-actin binding, point mutations there correlate with a loss of parasite motility and force generation (Kursula et al., 2008; Moreau et al., 2020; Nodelman et al., 1999; Qiao et al., 2019; Sun et al., 2018). Some structurally distinct eukaryotic Profilins form oligomers (dimers, trimers, and tetramers) that can reduce actin assembly mediated by Formins by reducing the affinity of Profilin for PLP and also obstructing Profilin-actin binding sites along Formin homology domains (Qiao et al., 2019; Sun et al., 2018). Profilin oligomers are also present in mammalian systems and are suggested to play important roles in disease (specifically neurodegeneration and immune responses), although the detailed mechanisms that underlie the physiological function and formation of these higher-order Profilin configurations are far from clear (Babich et al., 1996; Korupolu et al., 2009; Mares-Mejía et al., 2016; Posey et al., 2018).
Does a free Profilin pool exist in cells? Studies have consistently reported similar concentrations of actin and Profilin in diverse cell types and species. While these calculations are appealing to explain Profilin function with regard to actin monomers, they often fail to account for the presence of actin filaments or the affinity of Profilin for other ligands (i.e., Ena/VASP, Formins, activators of the Arp2/3 Complex, microtubules). Thus, with the assumption that 50–90% of cellular actin is polymerized into filaments (Funk et al., 2019; Pollard et al., 2000), at least half the total amount of Profilin may be available for functions beyond binding actin monomers or remaining unbound as a free Profilin pool. Does the presence of Profilin isoforms further complicate these interactions? Human Profilin isoforms 1–3 each bind actin, PIP lipids, and PLP albeit with different affinities, and whether the isoforms can be used interchangeably for these processes is unclear (Behnen et al., 2009; Lambrechts et al., 2006; Michaelsen et al., 2010). Profilin-2 binds PLP residues more strongly, but actin less efficiently, than Profilin-1 (Lu and Pollard, 2001; Vinson et al., 1998). Do these properties translate into more or less proficient Formin-mediated assembly? These interactions might slow Formin-mediated actin polymerization because this version of Profilin doesn’t bind actin well. On the other hand, these circumstances might aid Formin proteins in finding cytoplasmic actin monomers and more efficiently releasing Profilin from the polymerizing actin filament. Additionally, neuronal tissues are known to have more Profilin-2 than Profilin-1 and more tubulin than other tissues in the body (Witke et al., 1998). Although not explicitly tested, does this suggest that Profilin-2 is a better regulator of microtubule dynamics than Profilin-1? If so, this may further elucidate some of the complex details that underlie how cellular actin and microtubule dynamics are linked. Intriguingly, the testes-specific Profilin-4 isoform does not bind actin or PLP stretches, and mouse studies have demonstrated that each additional isoform of Profilin does not fully rescue the effects of Profilin-1. This suggests that there may be distinct tissue-specific roles for Profilin that go beyond actin assembly and define even more interactions that compete for cellular Profilin (Behnen et al., 2009; Polet et al., 2006; Witke et al., 1998).
Do we really understand the role of Profilin in cancer? Some have tried to bin specific Profilin isoforms as indicators of cancer prognosis with Profilin-1 behaving as a tumor suppressor and Profilin-2 suggesting malignancy, but there are multiple lines of evidence to contradict both of these statements. Data available from one of the most comprehensive databases quantifying RNA transcripts in cancers suggests that Profilin-1 and Profilin-2 RNA transcripts become elevated in most (but not all) cancers (Fig. 4). Is Profilin a good target for chemotherapies or to reinforce for better patient recovery? Regardless, to develop an effective pharmaceutical target we need to understand the exact timescales and mechanisms of Profilin function that are disrupted in disease and whether inhibiting one of them (i.e., actin binding) is enough. What if the role of Profilin goes beyond regulating cytoskeletal dynamics? Some reports suggest Profilin may play mysterious roles unrelated to the cytoskeleton in the nucleus (i.e., nuclear import/export and signaling), however if and how these studies link to cancer has not been elucidated (Holzinger et al., 2000; Lederer et al., 2005; Söderberg et al., 2012; Stüven et al., 2003). Alternatively, perhaps Profilin-facilitated post-translational modification of actin goes askew in cancer, ultimately leading to changes to specific forms of actin assembly and cell migratory behaviors (e.g., more actin N-terminal-acetylation decreases cell migration, and the formation of filopodia and lamellipodial protrusions) (Rebowski et al., 2020). Model organisms are powerful tools in this regard and can help to elucidate new and underexplored roles of Profilin in these and cancer-relevant cell processes.
Finally, an increasing number of recent studies link specific cancers directly to proteins that form or regulate the formation of biomolecular condensates (Chen et al., 2019; Kamagata et al., 2020). Although Profilin does not phase-separate on its own, it is found in many cellular condensates and can regulate their size and dynamics (Ghosh et al., 2019; Molliex et al., 2015; Posey et al., 2018). To date no studies have explored whether Profilin contributes to the pathological role of biomolecular condensates in cancer, however understanding exactly how Profilin regulates these processes could be a valuable asset in the development of cancer pharmaceutical agents and therapies.
ACKNOWLEDGMENTS
We are especially grateful to Marc R. Ridilla (Repair Biotechnologies), Amanda Young (SUNY Upstate), Naomi Courtemanche (University of Minnesota), Prachee Avasthi (University of Kansas Medical Center), Jonathan Wendel (Iowa State University), Fatima Cvrčková (Univerzita Karlova) for their helpful suggestions and/or for reading parts of this composition. This work was supported by the Michael E. Connolly Endowment for Lung Cancer Research and NIH 1R35GM133485-01 to J.L.H-R.
Footnotes
CONFLICTS OF INTEREST
The authors declare that they have no conflicts of interest.
REFERENCES
- Akhmanova A, Steinmetz MO, 2010. Microtubule +TIPs at a glance. J. Cell Sci 123, 3415–3419. [DOI] [PubMed] [Google Scholar]
- Akıl C, Robinson RC, 2018. Genomes of Asgard archaea encode profilins that regulate actin. Nature 562 (7727), 439–443. [DOI] [PubMed] [Google Scholar]
- Akin S, Can G, Durna Z, Aydiner A, 2008. The quality of life and self-efficacy of Turkish breast cancer patients undergoing chemotherapy. Eur. J. Oncol. Nurs 12 (5), 449–456. [DOI] [PubMed] [Google Scholar]
- Ali M, Heyob K, Jacob NK, Rogers LK, 2016. Alterative expression and localization of profilin-1/VASPpS157 and Cofilin-1/VASPpS239 regulates metastatic growth and is modified by DHA supplementation. Mol. Cancer Ther 15 (9), 2220–2231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alkam D, Feldman EZ, Singh A, Kiaei M, 2017. Profilin1 biology and its mutation, actin(g) in disease. Cell. Mol. Life Sci 74 (6), 967–981. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alushin GM, Lander GC, Kellogg EH, Zhang R, Baker D, Nogales E, 2014. High-resolution microtubule structures reveal the structural transitions in αβ-tubulin upon GTP hydrolysis. Cell 157 (5), 1117–1129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alvarado MI, Jimeno L, De La Torre F, Boissy P, Rivas B, Lázaro MJ, Barber D, 2014. Profilin as a severe food allergen in allergic patients overexposed to grass pollen. Allergy 69 (12), 1610–1616. [DOI] [PubMed] [Google Scholar]
- Amberg DC, Basart E, Botstein D, 1995. Defining protein interactions with yeast actin in vivo. Nat. Struct. Biol 2, 28–34. [DOI] [PubMed] [Google Scholar]
- Aroush DR-B, Ofer N, Abu-Shah E, Allard J, Krichevsky O, Mogilner A, Keren K, 2017. Actin turnover in lamellipodial fragments. Curr. Biol 27 (19), 2963–2973.e14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Augustine RC, Pattavina KA, Tüzel E, Vidali L, Bezanilla M, 2011. Actin interacting protein1 and actin depolymerizing factor drive rapid actin dynamics in Physcomitrella patens. Plant Cell 23, 3696–3710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aumeier C, Schaedel L, Gaillard J, John K, Blanchoin L, Théry M, 2016. Self-repair promotes microtubule rescue. Nat. Cell Biol 18 (10), 1054–1064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Avasthi P, Onishi M, Karpiak J, Yamamoto R, Mackinder L, Jonikas MC, Sale WS, Shoichet B, Pringle JR, Marshall WF, 2014. Actin is required for IFT regulation in Chlamydomonas reinhardtii. Curr. Biol 24 (17), 2025–2032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Babich M, Foti LRP, Sykaluk LL, Clark CR, 1996. Profilin forms tetramers that bind to G-actin. Biochem. Biophys. Res. Commun 218, 125–131. [DOI] [PubMed] [Google Scholar]
- Babik W, Dudek K, Fijarczyk A, Pabijan M, Stuglik M, Szkotak R, Zieliński P, 2014. Constraint and adaptation in newt toll-like receptor genes. Genome Biol. Evol 7 (1), 81–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bae YH, Ding Z, Zou L, Wells A, Gertler F, Roy P, 2009. Loss of profilin-1 expression enhances breast cancer cell motility by Ena/VASP proteins. J. Cell. Physiol 219 (2), 354–364. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bao Y, Hu G, Flagel LE, Salmon A, Bezanilla M, Paterson AH, Wang Z, Wendel JF, 2011. Parallel up-regulation of the profilin gene family following independent domestication of diploid and allopolyploid cotton (Gossypium). Proc. Natl. Acad. Sci 108 (52), 21152–21157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baraniskin A, Birkenkamp-Demtroder K, Maghnouj A, Zöllner H, Munding J, Klein-Scory S, Reinacher-Schick A, Schwarte-Waldhoff I, Schmiegel W, Hahn SA, 2012. MiR-30a-5p suppresses tumor growth in colon carcinoma by targeting DTL. Carcinogenesis 33 (4), 732–739. [DOI] [PubMed] [Google Scholar]
- Barshop BA, Wrenn RF, Frieden C, 1983. Analysis of numerical methods of computer simulation of kinetic processes: development of KINSIM—a flexible, portable system. Anal. Biochem 72, 248–254. [DOI] [PubMed] [Google Scholar]
- Baum J, Papenfuss AT, Baum B, Speed TP, Cowman AF, 2006. Regulation of apicomplexan actin-based motility. Nat. Rev. Microbiol 4 (8), 621–628. [DOI] [PubMed] [Google Scholar]
- Beck M, Schmidt A, Malmstroem J, Claassen M, Ori A, Szymborska A, Herzog F, Rinner O, Ellenberg J, Aebersold R, 2011. The quantitative proteome of a human cell line. Mol. Syst. Biol 7, 549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Behnen M, Murk K, Kursula P, Cappallo-Obermann H, Rothkegel M, Kierszenbaum AL, Kirchhoff C, 2009. Testis-expressed profilins 3 and 4 show distinct functional characteristics and localize in the acroplaxome-manchette complex in spermatids. BMC Cell Biol. 10, 34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bender M, Stritt S, Nurden P, van Eeuwijk JMM, Zieger B, Kentouche K, Schulze H, Morbach H, Stegner D, Heinze KG, Heinze K, Dütting S, Gupta S, Witke W, Falet H, Fischer A, Hartwig JH, Nieswandt B, 2014. Megakaryocyte-specific Profilin1-deficiency alters microtubule stability and causes a Wiskott-Aldrich syndrome-like platelet defect. Nat. Commun 5, 4746. [DOI] [PubMed] [Google Scholar]
- Berro J, Sirotkin V, Pollard TD, 2010. Mathematical modeling of endocytic actin patch kinetics in fission yeast: disassembly requires release of actin filament fragments. Mol. Biol. Cell 21, 2905–2915. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bertling E, Quintero-Monzon O, Mattila PK, Goode BL, Lappalainen P, 2007. Mechanism and biological role of profilin-Srv2/CAP interaction. J. Cell Sci 120, 1225–1234. [DOI] [PubMed] [Google Scholar]
- Blanchoin L, Pollard TD, 1998. Interaction of actin monomers with Acanthamoeba actophorin (ADF/cofilin) and profilin. J. Biol. Chem 273, 25106–25111. [DOI] [PubMed] [Google Scholar]
- Blanchoin L, Pollard TD, Mullins RD, 2000. Interactions of ADF/cofilin, Arp2/3 complex, capping protein and profilin in remodeling of branched actin filament networks. Curr. Biol 10, 1273–1282. [DOI] [PubMed] [Google Scholar]
- Blanchoin L, Boujemaa-Paterski R, Sykes C, Plastino J, 2014. Actin dynamics, architecture, and mechanics in cell motility. Physiol. Rev 94 (1), 235–263. [DOI] [PubMed] [Google Scholar]
- Blangy A, Lane HA, d’Hérin P, Harper M, Kress M, Nigg EA, 1995. Phosphorylation by p34cdc2 regulates spindle association of human Eg5, a kinesin-related motor essential for bipolar spindle formation in vivo. Cell 83 (7), 1159–1169. [DOI] [PubMed] [Google Scholar]
- Blasco R, Cole NB, Moss B, 1991. Sequence analysis, expression, and deletion of a vaccinia virus gene encoding a homolog of profilin, a eukaryotic actin-binding protein. J. Virol 65, 4598–4608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bolger-Munro M, Choi K, Scurll JM, Abraham L, Chappell RS, Sheen D, Dang-Lawson M, Wu X, Priatel JJ, Coombs D, Hammer JA, Gold MR, 2019. Arp2/3 complex-driven spatial patterning of the BCR enhances immune synapse formation, BCR signaling and B cell activation. eLife 8, e44574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bonello L, Camoin-Jau L, Armero S, Com O, Arques S, Burignat-Bonello C, Giacomoni M-P, Bonello R, Collet F, Rossi P, Barragan P, Dignat-George F, Paganelli F, 2009. Tailored clopidogrel loading dose according to platelet reactivity monitoring to prevent acute and subacute stent thrombosis. Am. J. Cardiol 103 (1), 5–10. [DOI] [PubMed] [Google Scholar]
- Boopathy S, Silvas TV, Tischbein M, Jansen S, Shandilya SM, Zitzewitz JA, Landers JE, Goode BL, Schiffer CA, Bosco DA, 2015. Structural basis for mutation-induced destabilization of profilin 1 in ALS. Proc. Natl. Acad. Sci 112 (26), 7984–7989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borisy G, Heald R, Howard J, Janke C, Musacchio A, Nogales E, 2016. Microtubules: 50 years on from the discovery of tubulin. Nat. Rev. Mol. Cell Biol 17 (5), 322–328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Böttcher RT, Wiesner S, Braun A, Wimmer R, Berna A, Elad N, Medalia O, Pfeifer A, Aszódi A, Costell M, Fässler R, 2009. Profilin 1 is required for abscission during late cytokinesis of chondrocytes. EMBO J. 28 (8), 1157–1169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bouchet BP, Akhmanova A, 2017. Microtubules in 3D cell motility. J. Cell Sci 130(1), 39–50. [DOI] [PubMed] [Google Scholar]
- Bouchet BP, Gough RE, Ammon Y-C, van de Willige D, Post H, Jacquemet G, Altelaar AM, Heck AJ, Goult BT, Akhmanova A, 2016. Talin-KANK1 interaction controls the recruitment of cortical microtubule stabilizing complexes to focal adhesions. Elife 5, e18124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brieher W, 2013. Mechanisms of actin disassembly. Mol. Biol. Cell 24 (15), 2299–2302. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bubb MR, Yarmola EG, Gibson BG, Southwick FS, 2003. Depolymerization of actin filaments by profilin: effects of profilin on capping protein function. J. Biol. Chem 278, 24629–24635. [DOI] [PubMed] [Google Scholar]
- Burbank KS, Mitchison TJ, 2006. Microtubule dynamic instability. Curr. Biol 16 (14), R516–R517. [DOI] [PubMed] [Google Scholar]
- Burke TA, Christensen JR, Barone E, Suarez C, Sirotkin V, Kovar DR, 2014. Homeostatic actin cytoskeleton networks are regulated by assembly factor competition for monomers. Curr. Biol 24 (5), 579–585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Burnett BG, Andrews J, Ranganathan S, Fischbeck KH, Di Prospero NA, 2008. Expression of expanded polyglutamine targets profilin for degradation and alters actin dynamics. Neurobiol. Dis 30 (3), 365–374. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Butler-Cole C, Wagner MJ, Da Silva M, Brown GD, Burke RD, Upton C, 2007. An ectromelia virus profilin homolog interacts with cellular tropomyosin and viral A-type inclusion protein. Virol. J 4, 76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Caglayan E, Romeo GR, Kappert K, Odenthal M, Südkamp M, Body SC, Shernan SK, Hackbusch D, Vantler M, Kazlauskas A, Rosenkranz S, 2010. Profilin-1 is expressed in human atherosclerotic plaques and induces atherogenic effects on vascular smooth muscle cells. PLoS One 5 (10), e13608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cande WZ, Lazarides E, McIntosh JR, 1977. A comparison of the distribution of actin and tubulin in the mammalian mitotic spindle as seen by indirect immunofluorescence. J. Cell Biol 72 (3), 552–567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cao L, Henty-Ridilla JL, Blanchoin L, Staiger CJ, 2016. Profilin-dependent nucleation and assembly of actin filaments controls cell elongation in Arabidopsis. Plant Physiol. 170 (1), 220–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Carlier M-F, Jean C, Rieger KJ, Lenfant M, Pantaloni D, 1993. Modulation of the interaction between g-actin and Thymosin-ß4 by ATP/ADP ratio: possible implication in the regulation of actin dynamics. Proc. Natl. Acad. Sci 90, 5034–5038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Carlier MF, Didry D, Erk I, Lepault J, Vantroys ML, Vandekerchkove J, Perelroizen I, Yin H, Doi YK, Pantaloni D, 1996. Tß4 is not a simple g-actin sequestering protein and interacts with F-actin at high concentration. J. Biol. Chem 271, 9231–9239. [DOI] [PubMed] [Google Scholar]
- Carlsson L, Nyström L-E, Lindberg U, 1976. Crystallization of a non-muscle actin. J. Mol. Biol 105, 353–366. [DOI] [PubMed] [Google Scholar]
- Carlsson L, Nyström LE, Sundkvist I, Markey F, Lindberg U, 1976. Profilin, a low-molecular weight protein controlling actin polymerisability. In: Contractile systems in non-muscle tissues, pp. 39–49. [DOI] [PubMed] [Google Scholar]
- Carlsson L, Nyström LE, Sundkvist I, Markey F, Lindberg U, 1977. Actin polymerizability is influenced by profilin, a low molecular weight protein in non-muscle cells. J. Mol. Biol 115 (3), 465–483. [DOI] [PubMed] [Google Scholar]
- Caudron N, Valiron O, Usson Y, Valiron P, Job D, 2000. A reassessment of the factors affecting microtubule assembly and disassembly in vitro. J. Mol. Biol 297 (1), 211–220. [DOI] [PubMed] [Google Scholar]
- Chaaban S, Brouhard GJ, 2017. A microtubule bestiary: structural diversity in tubulin polymers. Mol. Biol. Cell 28 (22), 2924–2931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chakraborty J, Pandey M, Navneet AK, Appukuttan TA, Varghese M, Sreetama SC, Rajamma U, Mohanakumar KP, 2014. Profilin-2 increased expression and its altered interaction with β-actin in the striatum of 3-nitropropionic acid-induced Huntington’s disease in rats. Neuroscience 281, 216–228. [DOI] [PubMed] [Google Scholar]
- Chan KY, Matthews KR, Ersfeld K, 2010. Functional characterisation and drug target validation of a mitotic Kinesin-13 in Trypanosoma brucei. PLoS Pathog. 6 (8), e1001050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang F, 2000. Microtubule and actin-dependent movement of the formin cdc12p in fission yeast. Microsc. Res. Tech 49, 161–167. [DOI] [PubMed] [Google Scholar]
- Chang F, Martin SG, 2009. Shaping fission yeast with microtubules. CSH Perspect.Biol 1 (1), a001347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang F, Drubin D, Nurse P, 1997. cdc12p, a protein required for cytokinesis in fission yeast, is a component of the cell division ring and interacts with profilin. J. Cell Biol 137, 169–182. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen RR, Yung MMH, Xuan Y, Zhan S, Leung LL, Liang RR, Leung THY, Yang H, Xu D, Sharma R, Chan KKL, Ngu S-F, Ngan HYS, Chan DW, 2019. Targeting of lipid metabolism with a metabolic inhibitor cocktail eradicates peritoneal metastases in ovarian cancer cells. Commun. Biol 2 (1), 281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chesarone MA, DuPage AG, Goode BL, 2010. Unleashing formins to remodel the actin and microtubule cytoskeletons. Nat. Rev. Mol. Cell Biol 11 (1), 62–74. [DOI] [PubMed] [Google Scholar]
- Chou SZ, Pollard TD, 2019. Mechanism of actin polymerization revealed by cryo-EM structures of actin filaments with three different bound nucleotides. Proc. Natl. Acad. Sci 116 (10), 4265–4274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Christensen JR, Craig EW, Glista MJ, Mueller DM, Li Y, Sees JA, Huang S, Suarez C, Mets LJ, Kovar DR, Avasthi P, 2019. Chlamydomonas reinhardtii formin FOR1 and profilin PRF1 are optimized FOR acute rapid actin filament assembly. Mol. Biol. Cell 30 (26), 3123–3135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coles CH, Bradke F, 2015. Coordinating neuronal actin-microtubule dynamics. Curr. Biol 25 (15), R677–R691. [DOI] [PubMed] [Google Scholar]
- Colin A, Singaravelu P, Théry M, Blanchoin L, Gueroui Z, 2018. Actin-network architecture regulates microtubule dynamics. Curr. Biol 28 (16), 2647–2656.e4. [DOI] [PubMed] [Google Scholar]
- Cooley L, Verheyen E, Ayers K, 1992. Chickadee encodes a profilin required for intercellular cytoplasm transport during Drosophila oogenesis. Cell 69 (1), 173–184. [DOI] [PubMed] [Google Scholar]
- Cooper JA, Walker SB, Pollard TD, 1983. Pyrene actin: documentation of the validity of a sensitive assay for actin polymerization. J. Muscle Res. Cell Motil 4, 253–262. [DOI] [PubMed] [Google Scholar]
- Cooper JA, Blum JD, Pollard TD, 1984. Acanthamoeba castellanii capping protein: properties, mechanism of action, immunologic cross-reactivity, and localization. J. Cell Biol 99, 217–225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coppolino MG, Dierckman R, Loijens J, Collins RF, Pouladi M, Jongstra-Bilen J, Schreiber AD, Trimble WS, Anderson R, Grinstein S, 2002. Inhibition of phosphatidylinositol-4-phosphate 5-kinase I alpha impairs localized actin remodeling and suppresses phagocytosis. J. Biol. Chem 277, 43849–43857. [DOI] [PubMed] [Google Scholar]
- Coumans JVF, Gau D, Poljak A, Wasinger V, Roy P, Moens PDJ, 2014. Profilin-1 overexpression in MDA-MB-231 breast cancer cells is associated with alterations in proteomics biomarkers of cell proliferation, survival, and motility as revealed by global proteomics analyses. Omics 18 (12), 778–791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Courtemanche N, 2018. Mechanisms of formin-mediated actin assembly and dynamics. Biophys. Rev 10 (6), 1553–1569. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Courtemanche N, Pollard TD, 2012. Determinants of formin homology 1 (FH1) domain function in actin filament elongation by formins. J. Biol. Chem 287 (10), 7812–7820. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Courtemanche N, Pollard TD, 2013. Interaction of profilin with the barbed end of actin filaments. Biochemistry 52 (37), 6456–6466. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Craig EW, Mueller DM, Bigge BM, Schaffer M, Engel BD, Avasthi P, 2019. The elusive actin cytoskeleton of a green alga expressing both conventional and divergent actins. Mol. Biol. Cell 30 (22), 2827–2837. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cui X, Zhang S, Xu Y, Dang H, Liu C, Wang L, Yang L, Hu J, Liang W, Jiang J, Li N, Li Y, Chen Y, Li F, 2016. PFN2, a novel marker of unfavorable prognosis, is a potential therapeutic target involved in esophageal squamous cell carcinoma.J. Transl. Med 14 (1), 137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davies T, Kim HX, Romano Spica N, Lesea-Pringle BJ, Dumont J, Shirasu-Hiza M, Canman JC, 2018. Cell-intrinsic and -extrinsic mechanisms promote cell-type-specific cytokinetic diversity. Elife 7, e36204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- De La Cruz EM, Ostap EM, Brundage RA, Reddy KS, Sweeney HL, Safer D, 2000. Thymosin-ß4 changes the conformation and dynamics of actin monomers. Biophys. J 78, 2516–2527. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Denkers EY, 2010. Toll-like receptor initiated host defense against Toxoplasma gondii.J. Biomed. Biotechnol 2010, 1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Desai A, Mitchison TJ, 1997. Microtubule polymerization dynamics. Annu. Rev. Cell Dev. Biol 13, 83–117. [DOI] [PubMed] [Google Scholar]
- Detmers PA, 1985. Elongation of cytoplasmic processes during gametic mating: models for actin-based motility. Can. J. Biochem. Cell Biol 63, 599–607. [DOI] [PubMed] [Google Scholar]
- Detmers PA, Goodenough UW, Condeelis J, 1983. Elongation of the fertilization tubule in Chlamydomonas: new observations on the core microfilaments and the effect of transient intracellular signals in their structural integrity. J. Cell Biol 97, 522–532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Di Nardo A, Gareus R, Kwiatkowski D, Witke W, 2000. Alternative splicing of the mouse profilin II gene generates functionally different profilin isoforms. J. Cell Sci 113, 3795–3803. [DOI] [PubMed] [Google Scholar]
- Didry D, Carlier M-F, Pantaloni D, 1998. Synergy between actin depolymerizing factor/cofilin and profilin in increasing actin filament turnover. J. Biol. Chem 273, 25602–25611. [DOI] [PubMed] [Google Scholar]
- Ding Z, Lambrechts A, Parepally M, Roy P, 2006. Silencing profilin-1 inhibits endothelial cell proliferation, migration and cord morphogenesis. J. Cell Sci 119, 4127–4137. [DOI] [PubMed] [Google Scholar]
- Ding Z, Joy M, Bhargava R, Gunsaulus M, Lakshman N, Miron-Mendoza M, Petroll M, Condeelis J, Wells A, Roy P, 2014. Profilin-1 downregulation has contrasting effects on early vs late steps of breast cancer metastasis. Oncogene 33 (16), 2065–2074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dogterom M, Koenderink GH, 2019. Actin-microtubule crosstalk in cell biology. Nat.Rev. Mol. Cell Biol 20 (1), 38–54. [DOI] [PubMed] [Google Scholar]
- Dominguez R, Holmes KC, 2011. Actin structure and function. Annu. Rev. Biophys 40, 169–186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dos Remedios CG, Chhabra D, Kekic M, Dedova IV, Tsubakihara M, Berry DA, Nosworthy NJ, 2003. Actin binding proteins: regulation of cytoskeletal microfilaments. Physiol. Rev 83, 433–473. [DOI] [PubMed] [Google Scholar]
- Dovas A, Gligorijevic B, Chen X, Entenberg D, Condeelis J, Cox D, 2011. Visualization of actin polymerization in invasive structures of macrophages and carcinoma cells using photoconvertible β-actin-Dendra2 fusion proteins. PLoS One 6 (2), e16485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dürre K, Keber FC, Bleicher P, Brauns F, Cyron CJ, Faix J, Bausch AR, 2018. Capping protein-controlled actin polymerization shapes lipid membranes. Nat. Commun 9 (1), 1630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dziezanowski MA, DeStefano MJ, Rabinovitch M, 1980. Effect of antitubulins on spontaneous and chemotactic migration of neutrophils under agarose. J. Cell Sci 42, 379–388. [DOI] [PubMed] [Google Scholar]
- Ehler E, 2018. Actin-associated proteins and cardiomyopathy—the ‘unknown’ beyond troponin and tropomyosin. Biophys. Rev 10 (4), 1121–1128. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Elie A, Prezel E, Guérin C, Denarier E, Ramirez-Rios S, Serre L, Andrieux A, Fourest-Lieuvin A, Blanchoin L, Arnal I, 2015. Tau co-organizes dynamic microtubule and actin networks. Sci. Rep 5, 9964. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Engelke MF, Winding M, Yue Y, Shastry S, Teloni F, Reddy S, Blasius TL, Soppina P, Hancock WO, Gelfand VI, Verhey KJ, 2016. Engineered kinesin motor proteins amenable to small-molecule inhibition. Nat. Commun 7 (1), 11159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Erickson HP, O’Brien ET, 1992. Microtubule dynamic instability and GTP hydrolysis. Annu. Rev. Biophys. Biomol. Struct 21 (1), 145–166. [DOI] [PubMed] [Google Scholar]
- Etienne-Manneville S, 2004. Actin and microtubules in cell motility: which one is in control?. Traffic 5 (7), 470–477. [DOI] [PubMed] [Google Scholar]
- Euteneuer U, Schliwa M, 1984. Persistent, directional motility of cells and cytoplasmic fragments in the absence of microtubules. Nature 310 (5972), 58–61. [DOI] [PubMed] [Google Scholar]
- Evangelista M, 1997. Bni1p, a yeast Formin linking Cdc42p and the actin cytoskeleton during polarized morphogenesis. Science 276 (5309), 118–122. [DOI] [PubMed] [Google Scholar]
- Evangelista M, Pruyne D, Amberg DC, Boone C, Bretscher A, 2002. Formins direct Arp2/3-independent actin filament assembly to polarize cell growth in yeast. Nat. Cell Biol 4 (3), 260–269. [DOI] [PubMed] [Google Scholar]
- Fabbri M, Wiemann J, Manucci F, Briggs DEG, 2020. Three-dimensional soft tissue preservation revealed in the skin of a non-avian dinosaur. Palaeontology 63 (2), 185–193. [Google Scholar]
- Fagerberg L, Hallström BM, Oksvold P, Kampf C, Djureinovic D, Odeberg J, Habuka M, Tahmasebpoor S, Danielsson A, Edlund K, Asplund A, Sjöstedt E, Lundberg E, Szigyarto CA-K, Skogs M, Takanen JO, Berling H, Tegel H, Mulder J, … Uhlén M, 2014. Analysis of the human tissue-specific expression by genome-wide integration of transcriptomics and antibody-based proteomics. Mol. Cell. Proteomics 13 (2), 397–406. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fan Y, Arif A, Gong Y, Jia J, Eswarappa SM, Willard B, Horowitz A, Graham LM, Penn MS, Fox PL, 2012. Stimulus-dependent phosphorylation of profilin-1 in angiogenesis. Nat. Cell Biol 14 (10), 1046–1056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fan Y, Potdar AA, Gong Y, Eswarappa SM, Donnola S, Lathia JD, Hambardzumyan D, Rich JN, Fox PL, 2014. Profilin-1 phosphorylation directs angiocrine expression and glioblastoma progression through HIF-1α accumulation. Nat. Cell Biol 16 (5), 445–456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Farina F, Gaillard J, Guérin C, Couté Y, Sillibourne J, Blanchoin L, Théry M, 2016. The centrosome is an actin-organizing centre, Nat. Cell Biol 18 (1), 65–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Farina F, Ramkumar N, Brown L, Samandar Eweis D, Anstatt J, Waring T, Bithell J, Scita G, Thery M, Blanchoin L, Zech T, Baum B, 2019. Local actin nucleation tunes centrosomal microtubule nucleation during passage through mitosis. EMBO J 38 (11). [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ferreira R, Limeta A, Nielsen J, 2019. Tackling cancer with yeast-based technologies. Trends Biotech. 37 (6), 592–603. [DOI] [PubMed] [Google Scholar]
- Ferron F, Rebowski G, Lee SH, Dominguez R, 2007. Structural basis for the recruitment of profilin-actin complexes during filament elongation by Ena/VASP. EMBO J. 26 (21), 4597–4606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fife CM, McCarroll JA, Kavallaris M, 2014. Movers and shakers: cell cytoskeleton in cancer metastasis. Brit. J. Pharm 171 (24), 5507–5523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Figley MD, Bieri G, Kolaitis R-M, Taylor JP, Gitler AD, 2014. Profilin 1 associates with stress granules and ALS-linked mutations alter stress granule dynamics. J. Neurosci 34 (24), 8083–8097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Frantzi M, van Kessel KE, Zwarthoff EC, Marquez M, Rava M, Malats N, Merseburger AS, Katafigiotis I, Stravodimos K, Mullen W, Zoidakis J, Makridakis M, Pejchinovski M, Critselis E, Lichtinghagen R, Brand K, Dakna M, Roubelakis MG, Theodorescu D, Vlahou A, Mischak H, Anagnou NP, 2016. Development and validation of urine-based peptide biomarker panels for detecting bladder cancer in a multi-center study. Clin. Cancer Res 22 (16), 4077–4086. [DOI] [PubMed] [Google Scholar]
- Fritzsche M, Fernandes RA, Chang VT, Colin-York H, Clausen MP, Felce JH, Galiani S, Erlenkämper C, Santos AM, Heddleston JM, Pedroza-Pacheco I, Waithe D, de la Serna JB, Lagerholm BC, Liu T, Chew T-L, Betzig E, Davis SJ, Eggeling C, 2017. Cytoskeletal actin dynamics shape a ramifying actin network underpinning immunological synapse formation. Sci. Adv 3 (6), e1603032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Funk J, Merino F, Venkova L, Heydenreich L, Kierfeld J, Vargas P, Raunser S, Piel M, Bieling P, 2019. Profilin and formin constitute a pacemaker system for robust actin filament growth. Elife 8, e50963. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fygenson DK, Flyvbjerg H, Sneppen K, Libchaber A, Leibler S, 1995. Spontaneous nucleation of microtubules. Phys. Rev. E 51 (5), 5058–5063. [DOI] [PubMed] [Google Scholar]
- Gaillard J, Ramabhadran V, Neumanne E, Gurel P, Blanchoin L, Vantard M, Higgs HN, 2011. Differential interactions of the formins INF2, mDia1, and mDia2 with microtubules. Mol. Biol. Cell 22 (23), 4575–4587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gardiner J, Marc J, 2011. Arabidopsis thaliana, a plant model organism for the neuronal microtubule cytoskeleton?. J. Exp. Bot 62 (1), 89–97. [DOI] [PubMed] [Google Scholar]
- Gareus R, Di Nardo A, Rybin V, Witke W, 2006. Mouse profilin 2 regulates endocytosis and competes with SH3 ligand binding to dynamin 1. J. Biol. Chem 281 (5), 2803–2811. [DOI] [PubMed] [Google Scholar]
- Gau D, Lewis T, McDermott L, Wipf P, Koes D, Roy P, 2018. Structure-based virtual screening identifies a small-molecule inhibitor of the profilin 1-actin interaction. J. Biol. Chem 293 (7), 2606–2616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gawadi N, 1971. The spindle at metaphase. Nature 232 (5305), 61–62. [DOI] [PubMed] [Google Scholar]
- Gertler FB, Niebuhr K, Reinhard M, Wehland J, Soriano P, 1996. Mena, a relative of VASP and Drosophila enabled, is implicated in the control of microfilament dynamics. Cell 87 (2), 227–239. [DOI] [PubMed] [Google Scholar]
- Ghiglione C, Jouandin P, Cérézo D, Noselli S, 2018. The Drosophila insulin pathway controls Profilin expression and dynamic actin-rich protrusions during collective cell migration. Development 145 (14), dev161117. [DOI] [PubMed] [Google Scholar]
- Ghosh A, Mazarakos K, Zhou H-X, 2019. Three archetypical classes of macromolecular regulators of protein liquid–liquid phase separation. Proc. Natl. Acad. Sci 116 (39), 19474–19483. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giansanti MG, Fuller MT, 2012. What Drosophila spermatocytes tell us about the mechanisms underlying cytokinesis. Cytoskeleton 69 (11), 869–881. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giansanti MG, Bonaccorsi S, Williams B, Williams EV, Santolamazza C, Goldberg ML, Gatti M, 1998. Cooperative interactions between the central spindle and the contractile ring during Drosophila cytokinesis. Genes Dev. 12 (3), 396–410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gieselmann R, Kwiatkowski DJ, Janmey PA, Witke W, 2008. Distinct biochemical characteristics of the two human profilin isoforms. Eur. J. Biol 229 (3), 621–628. [DOI] [PubMed] [Google Scholar]
- Goetz SC, Anderson KV, 2010. The primary cilium: a signaling centre during vertebrate development. Nat. Rev. Genet 11 (5), 331–344. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goldschmidt-Clermont PJ, Furman MI, Wachsstock D, Safer D, Nachmias VT, Pollard TD, 1992. The control of actin nucleotide exchange by thymosin beta-4 and profilin. A potential regulatory mechanism for actin polymerization in cells. Mol. Biol. Cell 3 (9), 1015–1024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Golemis EA, Scheet P, Beck TN, Scolnick EM, Hunter DJ, Hawk E, Hopkins N, 2018. Molecular mechanisms of the preventable causes of cancer in the United States. Genes Dev. 32 (13–14), 868–902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gordon JL, Sibley LD, 2005. Comparative genome analysis reveals a conserved family of actin-like proteins in apicomplexan parasites. BMC Genomics 6, 179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gordon DJ, Boyer JL, Korn ED, 1977. Comparative biochemistry of non-muscle actins. J. Biol. Chem 252 (22), 8300–8309. [PubMed] [Google Scholar]
- Grossman RL, Heath AP, Ferretti V, Varmus HE, Lowy DR, Kibbe WA, Staudt LM, 2016. Toward a shared vision for cancer genomic data. N. Engl. J. Med 375 (12), 1109–1112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haarer BK, Brown SS, 1990. Structure and function of profilin. Cell Motil. Cytoskeleton 17 (2), 71–74. [DOI] [PubMed] [Google Scholar]
- Hall A, 2009. The cytoskeleton and cancer. Cancer Metast. Rev 28 (1–2), 5–14. [DOI] [PubMed] [Google Scholar]
- Hancock WO, 2014. Mitotic kinesins: a reason to delve into Kinesin-12. Curr. Biol 24(19), R968–R970. [DOI] [PubMed] [Google Scholar]
- Hansson A, Skoglund G, Lassing I, Lindberg U, Ingelman-Sundberg M, 1988. Protein kinase C-dependent phosphorylation of profilin is specifically stimulated by phosphatidylinositol bisphosphate (PIP2). Biochem. Biophys. Res. Commun 150 (2), 526–531. [DOI] [PubMed] [Google Scholar]
- Harper JDI, McCurdy DW, Sanders MA, Salisbury JL, John PCL, 1992. Actin dynamics during the cell cycle in Chlamydomonas reinhardtii. Cell Motil. Cytoskeleton 22 (2), 117–126. [DOI] [PubMed] [Google Scholar]
- Harris EH, 2001. Chlamydomonas as a model organism. Annu. Rev. Plant. Physiol. Plant. Mol. Biol 52, 363–406. [DOI] [PubMed] [Google Scholar]
- Hauser M, Roulias A, Ferreira F, Egger M, 2010. Panallergens and their impact on the allergic patient. Allergy Asthma Clin. Immunol 6 (1), 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henty-Ridilla JL, Goode BL, 2015. Global resource distribution: allocation of actin building blocks by profilin. Dev. Cell 32 (1), 5–6. [DOI] [PubMed] [Google Scholar]
- Henty-Ridilla JL, Shimono M, Li J, Chang JH, Day B, Staiger CJ, 2013. The plant actin cytoskeleton responds to signals from microbe-associated molecular patterns. PLoS Pathog. 9 (4), e1003290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henty-Ridilla JL, Rankova A, Eskin JA, Kenny K, Goode BL, 2016. Accelerated actin filament polymerization from microtubule plus ends. Science 352 (6288), 1004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henty-Ridilla JL, Juanes MA, Goode BL, 2017. Profilin directly promotes microtubule growth through residues mutated in amyotrophic lateral sclerosis. Curr. Biol 27 (22), 3535–3543.e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Herman IM, Pollard TD, 1979. Comparison of purified anti-actin and fluorescent-heavy meromyosin staining patterns in dividing cells. J. Cell Biol 80 (3), 509–520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hernandez P, Tirnauer JS, 2010. Tumor suppressor interactions with microtubules: keeping cell polarity and cell division on track. Dis. Model Mech 3 (5–6), 304–315. [DOI] [PubMed] [Google Scholar]
- Hetrick B, Han MS, Helgeson LA, Nolen BJ, 2013. Small molecules CK-666 and CK-869 inhibit actin-related protein 2/3 complex by blocking an activating conformational change. Chem. Biol 20 (5), 701–712. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Higgins M, Obaidi I, McMorrow T, 2019. Primary cilia and their role in cancer. Oncol. Lett 17 (3), 3041–3047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Higgs HN, Pollard TD, 1999. Regulation of actin polymerization by Arp2/3 complex and WASp/scar proteins. J. Biol. Chem 274 (46), 32531–32534. [DOI] [PubMed] [Google Scholar]
- Holzinger A, Valenta R, Lütz-Meindl U, 2000. Profilin is localized in the nucleus-associated microtubule and actin system and is evenly distributed in the cytoplasm of the green algaMicrasterias denticulata. Protoplasma 212 (3), 197–205. [Google Scholar]
- Horan BG, Zerze GH, Kim YC, Vavylonis D, Mittal J, 2018. Computational modeling highlights the role of the disordered formin homology 1 domain in profilin-actin transfer. FEBS Lett. 592 (11), 1804–1816. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Horrevoets AJG, 2007. Profilin-1: an unexpected molecule linking vascular inflammation to the actin cytoskeleton. Circ. Res 101 (4), 328–330. [DOI] [PubMed] [Google Scholar]
- Huang B, Wang W, Bates M, Zhuang X, 2008. Three-dimensional super-resolution imaging by stochastic optical reconstruction microscopy. Science 319 (5864), 810–813. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hurst DR, Welch DR, 2011. Metastasis suppressor genes at the interface between the environment and tumor cell growth. Int. Rev. Cell Mol. Biol 286, 107–180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Inoue D, Obino D, Pineau J, Farina F, Gaillard J, Guerin C, Blanchoin L, Lennon-Duménil A-M, Théry M, 2019. Actin filaments regulate microtubule growth at the centrosome. EMBO J. 38 (11). [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ishizaki T, Morishima Y, Okamoto M, Furuyashiki T, Kato T, Narumiya S, 2001. Coordination of microtubules and the actin cytoskeleton by the rho effector mDia1. Nat. Cell Biol 3 (1), 8–14. [DOI] [PubMed] [Google Scholar]
- Isogai T, van der Kammen R, Innocenti M, 2015. SMIFH2 has effects on Formins and p53 that perturb the cell cytoskeleton. Sci. Rep 5, 9802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jack B, Mueller DM, Fee AC, Tetlow AL, Avasthi P, 2019. Partially redundant actin genes in Chlamydomonas control transition zone organization and flagellum-directed traffic. Cell Rep. 27 (8), 2459–2467.e3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Janke J, Schlüter K, Jandrig B, Theile M, Kölble K, Arnold W, Grinstein E, Schwartz A, Estevéz-Schwarz L, Schlag PM, Jockusch BM, Scherneck S, 2000. Suppression of tumorigenicity in breast cancer cells by the microfilament protein profilin 1. J. Exp. Med 191 (10), 1675–1686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jégou A, Niedermayer T, Orbán J, Didry D, Lipowsky R, Carlier M-F, Romet-Lemonne G, 2011. Individual actin filaments in a microfluidic flow reveal the mechanism of ATP hydrolysis and give insight into the properties of profilin. PLoS Biol. 9 (9), e1001161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang H, Wang S, Huang Y, He X, Cui H, Zhu X, Zheng Y, 2015. Phase transition of spindle-associated protein regulate spindle apparatus assembly. Cell 163 (1), 108–122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang C, Ding Z, Joy M, Chakraborty S, Kim SH, Bottcher R, Condeelis J, Singh S, Roy P, 2017. A balanced level of profilin-1 promotes stemness and tumor-initiating potential of breast cancer cells. Cell Cycle 16 (24), 2366–2373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Job D, Valiron O, Oakley B, 2003. Microtubule nucleation. Curr. Opin. Cell Biol 15 (1), 111–117. [DOI] [PubMed] [Google Scholar]
- Joy ME, Vollmer LL, Hulkower K, Stern AM, Peterson CK, Boltz RCD, Roy P, Vogt A, 2014. A high-content, multiplexed screen in human breast cancer cells identifies profilin-1 inducers with anti-migratory activities. PLoS One 9 (2), e88350. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Juanes MA, Bouguenina H, Eskin JA, Jaiswal R, Badache A, Goode BL, 2017. Adenomatous polyposis coli nucleates actin assembly to drive cell migration and microtubule-induced focal adhesion turnover. J. Cell Biol 216 (9), 2859–2875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Juanes MA, Isnardon D, Badache A, Brasselet S, Mavrakis M, Goode BL, 2019. The role of APC-mediated actin assembly in microtubule capture and focal adhesion turnover. J. Cell Biol 218 (10), 3415–3435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaiser DA, Vinson VK, Murphy DB, Pollard TD, 1999. Profilin is predominantly associated with monomeric actin in Acanthamoeba. J. Cell Sci 112, 3779–3790. [DOI] [PubMed] [Google Scholar]
- Kamagata K, Kanbayashi S, Honda M, Itoh Y, Takahashi H, Kameda T, Nagatsugi F, Takahashi S, 2020. Liquid-like droplet formation by tumor suppressor p53 induced by multivalent electrostatic interactions between two disordered domains. Sci. Rep 10 (1), 580. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kandasamy MK, McKinney EC, Meagher RB, 2002. Plant profilin isovariants are distinctly regulated in vegetative and reproductive tissues. Cell Motil. Cytoskeleton 52 (1), 22–32. [DOI] [PubMed] [Google Scholar]
- Kaverina I, Straube A, 2011. Regulation of cell migration by dynamic microtubules. Semin. Cell Dev. Biol 22 (9), 968–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaverina I, Rottner K, Small JV, 1998. Targeting, capture, and stabilization of microtubules at early focal adhesions. J. Cell Biol 142 (1), 181–190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim J-G, Moon M-Y, Kim H-J, Li Y, Song D-K, Kim J-S, Lee J-Y, Kim J, Kim S-C, Park J-B, 2012. Ras-related GTPases Rap1 and RhoA collectively induce the phagocytosis of serum-opsonized zymosan particles in macrophages. J. Biol. Chem 287 (7), 5145–5155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim M-J, Lee Y-S, Han G-Y, Lee H-N, Ahn C, Kim C-W, 2015. Profilin 2 promotes migration, invasion, and stemness of HT29 human colorectal cancer stem cells. Biosci. Biotechnol. Biochem 79 (9), 1438–1446. [DOI] [PubMed] [Google Scholar]
- Kita AM, Swider ZT, Erofeev I, Halloran MC, Goryachev AB, Bement WM, 2019. Spindle-F-actin interactions in mitotic spindles in an intact vertebrate epithelium. Mol. Biol. Cell 30 (14), 1645–1654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kollman JM, Polka JK, Zelter A, Davis TN, Agard DA, 2010. Microtubule nucleating gamma-TuSC assembles structures with 13-fold microtubule-like symmetry. Nature 466 (7308), 879–882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kooij V, Viswanathan MC, Lee DI, Rainer PP, Schmidt W, Kronert WA, Harding SE, Kass DA, Bernstein SI, Van Eyk JE, Cammarato A, 2016. Profilin modulates sarcomeric organization and mediates cardiomyocyte hypertrophy. Cardiovasc. Res 110 (2), 238–248. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Korupolu RV, Achary MS, Aneesa F, Sathish K, Wasia R, Sairam M, Nagarajaram HA, Singh SS, 2009. Profilin oligomerization and its effect on poly(L-)proline binding and phosphorylation. Int. J. Biol. Macromol 45 (3), 265–273. [DOI] [PubMed] [Google Scholar]
- Kotila T, Kogan K, Enkavi G, Guo S, Vattulainen I, Goode BL, Lappalainen P, 2018. Structural basis of actin monomer re-charging by cyclase-associated protein. Nat. Commun 9 (1), 1892. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kovar DR, Drøbak BK, Staiger CJ, 2000. Maize profilin isoforms are functionally distinct. Plant Cell 12 (4), 583–598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kovar DR, Yang P, Sale WS, Drobak BK, Staiger CJ, 2001. Chlamydomonas reinhardtii produces a profilin with unusual biochemical properties. J. Cell Sci 114 (23), 4293–4305. [DOI] [PubMed] [Google Scholar]
- Kovar DR, Kuhn JR, Tichy AL, Pollard TD, 2003. The fission yeast cytokinesis formin Cdc12p is a barbed end actin filament capping protein gated by profilin. J. Cell Biol 161 (5), 875–887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kovar DR, Harris ES, Mahaffy R, Higgs HN, Pollard TD, 2006. Control of the assembly of ATP- and ADP-actin by formins and profilin. Cell 124 (2), 423–435. [DOI] [PubMed] [Google Scholar]
- Krause M, Dent EW, Bear JE, Loureiro JJ, Gertler FB, 2003. Ena/VASP proteins: regulators of the actin cytoskeleton and cell migration. Annu. Rev. Cell Dev. Biol 19, 541–564. [DOI] [PubMed] [Google Scholar]
- Krishnan K, Moens PDJ, 2009. Structure and functions of profilins. Biophys. Rev 1 (2), 71–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kucera K, Koblansky AA, Saunders LP, Frederick KB, De La Cruz EM, Ghosh S, Modis Y, 2010. Structure-based analysis of Toxoplasma gondii profilin: a parasite-specific motif is required for recognition by toll-like receptor 11. J. Mol. Biol 403 (4), 616–629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kursula P, Kursula I, Massimi M, Song Y-H, Downer J, Stanley WA, Witke W, Wilmanns M, 2008. High-resolution structural analysis of mammalian profilin 2a complex formation with two physiological ligands: the formin homology 1 domain of mDia1 and the proline-rich domain of VASP. J. Mol. Biol 375 (1), 270–290. [DOI] [PubMed] [Google Scholar]
- Kwak J, Park OK, Jung YJ, Hwang BJ, Kwon S-H, Kee Y, 2013. Live image profiling of neural crest lineages in zebrafish transgenic lines. Mol. Cells 35 (3), 255–260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lai S-L, Chan T-H, Lin M-J, Huang W-P, Lou S-W, Lee S-J, 2008. Diaphanous-related formin 2 and profilin I are required for gastrulation cell movements. PLoS One 3 (10), e3439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lambrechts A, Verschelde JL, Jonckheere V, Goethals M, Vandekerckhove J, Ampe C, 1997. The mammalian profilin isoforms display complementary affinities for PIP2 and proline-rich sequences. EMBO J. 16 (3), 484–494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lambrechts A, Jonckheere V, Peleman C, Polet D, De Vos W, Vandekerckhove J, Ampe C, 2006. Profilin-I-ligand interactions influence various aspects of neuronal differentiation. J. Cell Sci 119 (8), 1570–1578. [DOI] [PubMed] [Google Scholar]
- Lämmermann T, Sixt M, 2009. Mechanical modes of “amoeboid” cell migration. Curr. Opin. Cell Biol 21 (5), 636–644. [DOI] [PubMed] [Google Scholar]
- Lassing I, Lindberg U, 1985. Specific interaction between phosphatidylinositol 4,5-bisphosphate and profilactin. Nature 314 (6010), 472–474. [DOI] [PubMed] [Google Scholar]
- LeCorgne H, Tudosie AM, Lavik K, Su R, Becker KN, Moore S, Walia Y, Wisner A, Koehler D, Alberts AS, Williams FE, Eisenmann KM, 2018. Differential toxicity of mDia formin-directed functional agonists and antagonists in developing zebrafish. Front. Pharmacol 9, 340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lederer M, Jockusch BM, Rothkegel M, 2005. Profilin regulates the activity of p42POP, a novel Myb-related transcription factor. J. Cell Sci 118 (2), 331–341. [DOI] [PubMed] [Google Scholar]
- Lee Y-J, Mazzatti DJ, Yun Z, Keng PC, 2005. Inhibition of invasiveness of human lung cancer cell line H1299 by over-expression of cofilin. Cell Biol. Int 29 (11), 877–883. [DOI] [PubMed] [Google Scholar]
- Lewkowicz E, Herit F, Le Clainche C, Bourdoncle P, Perez F, Niedergang F, 2008. The microtubule-binding protein CLIP-170 coordinates mDia1 and actin reorganization during CR3-mediated phagocytosis. J. Cell Biol 183 (7), 1287–1298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li R, Gundersen GG, 2008. Beyond polymer polarity: how the cytoskeleton builds a polarized cell. Nat. Rev. Mol. Cell Biol 9 (11), 860–873. [DOI] [PubMed] [Google Scholar]
- Lila T, Drubin DG, 1997. Evidence for physical and functional interactions among two Saccharomyces cerevisiae SH3 domain proteins, an adenylyl cyclase-associated protein and the actin cytoskeleton. Mol. Biol. Cell 8 (2), 367–385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lodish H, Berk A, Zipursky SL, et al. , 2000. The actin cytoskeleton In: Freeman WH (Ed.), Molecular Cell Biology. Macmillan Education, New York, NY, USA. [Google Scholar]
- Loisel TP, Boujemaa R, Pantaloni D, Carlier MF, 1999. Reconstitution of actin-based motility of Listeria and Shigella using pure proteins. Nature 401 (6753), 613–616. [DOI] [PubMed] [Google Scholar]
- Löwe J, Li H, Downing KH, Nogales E, 2001. Refined structure of alpha beta-tubulin at 3.5 Å resolution. J. Mol. Biol 313 (5), 1045–1057. [DOI] [PubMed] [Google Scholar]
- Lu J, Pollard TD, 2001. Profilin binding to poly-L-proline and actin monomers along with ability to catalyze actin nucleotide exchange is required for viability of fission yeast. Mol. Biol. Cell 12 (4), 1161–1175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Machesky LM, Pollard TD, 1993. Profilin as a potential mediator of membrane-cytoskeleton communication. Trends Cell Biol. 3 (11), 381–385. [DOI] [PubMed] [Google Scholar]
- Machesky LM, Cole NB, Moss B, Pollard TD, 1994. Vaccinia virus expresses a novel profilin with a higher affinity for polyphosphoinositides than actin. Biochemistry 33 (35), 10815–10824. [DOI] [PubMed] [Google Scholar]
- Machesky LM, Mullins RD, Higgs HN, Kaiser DA, Blanchoin L, May RC, Hall ME, Pollard TD, 1999. Scar, a WASp-related protein, activates nucleation of actin filaments by the Arp2/3 complex. Proc. Natl. Acad. Sci 96 (7), 3739–3744. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Majesky MW, 2007. Developmental basis of vascular smooth muscle diversity. Arterioscler. Thromb. Vasc. Biol 27 (6), 1248–1258. [DOI] [PubMed] [Google Scholar]
- Mammoto A, Sasaki T, Asakura T, Hotta I, Imamura H, Takahashi K, Matsuura Y, Shirao T, Takai Y, 1998. Interactions of drebrin and gephyrin with profilin. Biochem. Biophys. Res. Commun 243 (1), 86–89. [DOI] [PubMed] [Google Scholar]
- Manandhar A, Kang M, Chakraborty K, Loverde SM, 2018. Effect of nucleotide state on the protofilament conformation of tubulin octamers. J. Phys. Chem 122 (23), 6164–6178. [DOI] [PubMed] [Google Scholar]
- Mares-Mejía I, Martínez-Caballero S, Garay-Canales C, Cano-Sánchez P, Torres-Larios A, Lara-González S, Ortega E, Rodríguez-Romero A, 2016. Structural insights into the IgE mediated responses induced by the allergens Hev b 8 and Zea m 12 in their dimeric forms. Sci. Rep 6, 32552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Margolis RL, 1981. Role of GTP hydrolysis in microtubule treadmilling and assembly. Proc. Natl. Acad. Sci 78 (3), 1586–1590. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McGrath JL, Tardy Y, Dewey CF, Meister JJ, Hartwig JH, 1998. Simultaneous measurements of actin filament turnover, filament fraction, and monomer diffusion in endothelial cells. Biophys. J 75 (4), 2070–2078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McIntosh JR, Hays T, 2016. A brief history of research on mitotic mechanisms. Biology 5 (4), 55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Melak M, Plessner M, Grosse R, 2017. Actin visualization at a glance. J. Cell Sci 130 (3), 525–530. [DOI] [PubMed] [Google Scholar]
- Melamed Z, López-Erauskin J, Baughn MW, Zhang O, Drenner K, Sun Y, Freyermuth F, McMahon MA, Beccari MS, Artates JW, Ohkubo T, Rodriguez M, Lin N, Wu D, Bennett CF, Rigo F, Da Cruz S, Ravits J, Lagier-Tourenne C, Cleveland DW, 2019. Premature polyadenylation-mediated loss of stathmin-2 is a hallmark of TDP-43-dependent neurodegeneration. Nat. Neurosci 22 (2), 180–190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Merino F, Pospich S, Funk J, Wagner T, Küllmer F, Arndt H-D, Bieling P, Raunser S, 2018. Structural transitions of F-actin upon ATP hydrolysis at near-atomic resolution revealed by cryo-EM. Nat. Struct. Mol. Biol 25 (6), 528–537. [DOI] [PubMed] [Google Scholar]
- Meyers JR, 2018. Zebrafish: development of a vertebrate model organism. Curr. Protoc. Lab. Tech 16 (1), e19. [Google Scholar]
- Michaelsen K, Murk K, Zagrebelsky M, Dreznjak A, Jockusch BM, Rothkegel M, Korte M, 2010. Fine-tuning of neuronal architecture requires two profilin isoforms. Proc. Natl. Acad. Sci 107 (36), 15780–15785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Michelot A, Drubin DG, 2011. Building distinct actin filament networks in a common cytoplasm. Curr. Biol 21 (14), R560–R569. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miki H, Suetsugu S, Takenawa T, 1998. WAVE, a novel WASP-family protein involved in actin reorganization induced by Rac. EMBO J. 17 (23), 6932–6941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Milic B, Chakraborty A, Han K, Bassik MC, Block SM, 2018. KIF15 nanomechanics and kinesin inhibitors, with implications for cancer chemotherapeutics. Proc. Natl. Acad. Sci 115 (20), E4613–E4622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miller AL, 2011. The contractile ring. Curr. Biol 21 (24), R976–R978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mitchison T, 1993. Localization of an exchangeable GTP binding site at the plus end of microtubules. Science 261 (5124), 1044–1047. [DOI] [PubMed] [Google Scholar]
- Mitchison T, Kirschner M, 1984. Microtubule assembly nucleated by isolated centrosomes. Nature 312 (5991), 232–237. [DOI] [PubMed] [Google Scholar]
- Mitchison T, Kirschner M, 1984. Dynamic instability of microtubule growth. Nature 312 (5991), 237–242. [DOI] [PubMed] [Google Scholar]
- Mockrin SC, Korn ED, 1980. Acanthamoeba profilin interacts with G-actin to increase the rate of exchange of actin-bound adenosine 5′-triphosphate. Biochem. 19 (23), 5359–5362. [DOI] [PubMed] [Google Scholar]
- Mogilner A, Oster G, 1996. Cell motility driven by actin polymerization. Biophys. J 71 (6), 3030–3045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Molliex A, Temirov J, Lee J, Coughlin M, Kanagaraj AP, Kim HJ, Mittag T, Taylor JP, 2015. Phase separation by low complexity domains promotes stress granule assembly and drives pathological fibrillization. Cell 163 (1), 123–133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moreau CA, Bhargav SP, Kumar H, Quadt KA, Piirainen H, Strauss L, Kehrer J, Streichfuss M, Spatz JP, Wade RC, Kursula I, Frischknecht F, 2017. A unique profilin-actin interface is important for malaria parasite motility. PLOS Pathog. 13 (5), e1006412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moreau CA, Quadt KA, Piirainen H, Kumar H, Bhargav SP, Strauss L, Tolia NH, Wade RC, Spatz JP, Kursula I, Frischknecht F. 2020. A function of profilin in force generation during malaria parasite motility independent of actin binding. J. Cell Sci 134 (5), jcs233775 (2020). [DOI] [PubMed] [Google Scholar]
- Moritz M, Braunfeld MB, Guénebaut V, Heuser J, Agard DA, 2000. Structure of the gamma-tubulin ring complex: a template for microtubule nucleation. Nat. Cell Biol 2 (6), 365–370. [DOI] [PubMed] [Google Scholar]
- Mostowy S, Shenoy AR, 2015. The cytoskeleton in cell-autonomous immunity: structural determinants of host defense. Nat. Rev. Immunol 15 (9), 559–573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mouneimne G, Hansen SD, Selfors LM, Petrak L, Hickey MM, Gallegos LL, Simpson KJ, Lim J, Gertler FB, Hartwig JH, Mullins RD, Brugge JS, 2012. Differential remodeling of actin cytoskeleton architecture by profilin isoforms leads to distinct effects on cell migration and invasion. Cancer Cell 22 (5), 615–630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moustakas A, Heldin C-H, 2008. Dynamic control of TGF-β signaling and its links to the cytoskeleton. FEBS Lett. 582 (14), 2051–2065. [DOI] [PubMed] [Google Scholar]
- Mukhtar E, Adhami VM, Mukhtar H, 2014. Targeting microtubules by natural agents for cancer therapy. Mol. Cancer Ther 13 (2), 275–284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mullins RD, Pollard TD, 1999. Structure and function of the Arp2/3 complex. Curr.Opin. Struct. Biol 9 (2), 244–249. [DOI] [PubMed] [Google Scholar]
- Mullins RD, Heuser JA, Pollard TD, 1998. The interaction of Arp2/3 complex with actin: nucleation, high affinity pointed end capping, and formation of branching networks of filaments. Proc. Natl. Acad. Sci 95 (11), 6181–6186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murk K, Buchmeier S, Jockusch BM, Rothkegel M, 2009. In birds, profilin-2a is ubiquitously expressed and contributes to actin-based motility. J. Cell Sci 122 (7), 957–964. [DOI] [PubMed] [Google Scholar]
- Müssar KJ, Kandasamy MK, McKinney EC, Meagher RB, 2015. Arabidopsis plants deficient in constitutive class profilins reveal independent and quantitative genetic effects. BMC Plant Biol. 15, 177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Neidt EM, Scott BJ, Kovar DR, 2009. Formin differentially utilizes profilin isoforms to rapidly assemble actin filaments. J. Biol. Chem 284 (1), 673–684. [DOI] [PubMed] [Google Scholar]
- Nejedla M, Sadi S, Sulimenko V, de Almeida FN, Blom H, Draber P, Aspenström P, Karlsson R, 2016. Profilin connects actin assembly with microtubule dynamics. Mol. Biol. Cell 27 (15), 2381–2393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nejedla M, Li Z, Masser AE, Biancospino M, Spiess M, Mackowiak SD, Friedländer MR, Karlsson R, 2017. A fluorophore fusion construct of human profilin-I with non-compromised poly(L-)proline binding capacity suitable for imaging. J. Mol. Biol 429 (7), 964–976. [DOI] [PubMed] [Google Scholar]
- Nodelman IM, Bowman GD, Lindberg U, Schutt CE, 1999. X-ray structure determination of human profilin II: a comparative structural analysis of human profilins. J. Mol. Biol 294 (5), 1271–1285. [DOI] [PubMed] [Google Scholar]
- Nogales E, 2001. Structural insight into microtubule function. Annu. Rev. Biophys. Biomol. Struct 30, 397–420. [DOI] [PubMed] [Google Scholar]
- Nogales E, Wolf SG, Downing KH, 1998. Structure of the alpha beta tubulin dimer by electron crystallography. Nature 391 (6663), 199–203. [DOI] [PubMed] [Google Scholar]
- Nolen BJ, Tomasevic N, Russell A, Pierce DW, Jia Z, McCormick CD, Hartman J, Sakowicz R, Pollard TD, 2009. Characterization of two classes of small molecule inhibitors of Arp2/3 complex. Nature 460 (7258), 1031–1034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oakley BR, Paolillo V, Zheng Y, 2015. γ-Tubulin complexes in microtubule nucleation and beyond. Mol. Biol. Cell 26 (17), 2957–2962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oda T, Aihara T, Wakabayashi K, 2016. Early nucleation events in the polymerization of actin, probed by time-resolved small-angle x-ray scattering. Sci. Rep 6, 34539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Onishi M, Pringle JR, Cross FR, 2016. Evidence that an unconventional actin can provide essential F-actin function and that a surveillance system monitors F-actin integrity in Chlamydomonas. Genetics 202 (3), 977–996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ono S, 2013. The role of cyclase-associated protein in regulating actin filament dynamics - more than a monomer-sequestration factor. J. Cell Sci 126 (15), 3249–3258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ostrander DB, Ernst EG, Lavoie TB, Gorman JA, 1999. Polyproline binding is an essential function of human profilin in yeast. Eur. J. Biochem 262 (1), 26–35. [DOI] [PubMed] [Google Scholar]
- Palucka AK, Coussens LM, 2016. The basis of oncoimmunology. Cell 164 (6),1233–1247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pan L-N, Zhang Y, Zhu C-J, Dong Z-X, 2017. Kinesin KIF4A is associated with chemotherapeutic drug resistance by regulating intracellular trafficking of lung resistance-related protein. J. Zhejiang Univ. Sci. B 18 (12), 1046–1054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pantaloni D, Carlier MF, 1993. How profilin promotes actin filament assembly in the presence of Thymosin Beta 4. Cell 75 (5), 1007–1014. [DOI] [PubMed] [Google Scholar]
- Pantaloni D, Le Clainche C, Carlier MF, 2001. Mechanism of actin-based motility. Science 292 (5521), 1502–1506. [DOI] [PubMed] [Google Scholar]
- Panzica MT, Marin HC, Reymann A-C, McNally FJ, 2017. F-actin prevents interaction between sperm DNA and the oocyte meiotic spindle in C. elegans. J. Cell Biol 216 (8), 2273–2282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paplomata E, O’Regan R, 2014. The PI3K/AKT/mTOR pathway in breast cancer: targets, trials and biomarkers. Ther. Adv. Med. Oncol 6 (4), 154–166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paul A, Pollard T, 2008. The role of the FH1 domain and profilin in formin-mediated actin-filament elongation and nucleation. Curr. Biol 18 (1), 9–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pazour GJ, Witman GB, 2009. The Chlamydomonas flagellum as a model for human ciliary disease In: Chlamydomonas Sourcebook. Academy Press, pp. 445–478. [Google Scholar]
- Pelham RJ, Chang F, 2002. Actin dynamics in the contractile ring during cytokinesis in fission yeast. Nature 419 (6902), 82–86. [DOI] [PubMed] [Google Scholar]
- Percipalle P, Vartiainen M, 2019. Cytoskeletal proteins in the cell nucleus: a special nuclear actin perspective. Mol. Biol. Cell 30 (15), 1781–1785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perelroizen I, Marchand J-B, Blanchoin L, Didry D, Carlier M-F, 1994. Interaction of profilin with G-actin and poly(L-)proline. Biochemistry 33 (28), 8472–8478. [DOI] [PubMed] [Google Scholar]
- Pernier J, Shekhar S, Jegou A, Guichard B, Carlier M-F, 2016. Profilin interaction with actin filament barbed end controls dynamic instability, capping, branching, and motility. Dev. Cell 36 (2), 201–214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peterson JR, Bickford LC, Morgan D, Kim AS, Ouerfelli O, Kirschner MW, Rosen MK, 2004. Chemical inhibition of N-WASP by stabilization of a native autoinhibited conformation. Nat. Struct. Mol. Biol 11 (8), 747–755. [DOI] [PubMed] [Google Scholar]
- Petrella EC, Machesky LM, Kaiser DA, Pollard TD, 1996. Structural requirements and thermodynamics of the interaction of proline peptides with profilin. Biochemistry 35 (51), 16535–16543. [DOI] [PubMed] [Google Scholar]
- Pfajfer L, Mair NK, Jiménez-Heredia R, Genel F, Gulez N, Ardeniz, Hoeger B, Bal SK, Madritsch C, Kalinichenko A, Chandra Ardy R, Gerçeker B, Rey-Barroso J, Ijspeert H, Tangye SG, Simonitsch-Klupp I, Huppa JB, van der Burg M, Dupré L, Boztug K, 2018. Mutations affecting the actin regulator WD repeat-containing protein 1 lead to aberrant lymphoid immunity. J. Allergy Clin. Immunol. Pract 142 (5), 1589–1604.e11. [DOI] [PubMed] [Google Scholar]
- Pinto-Costa R, Sousa MM, 2019. Profilin as a dual regulator of actin and microtubule dynamics. Cytoskeleton 77 (3–4), 76–83. [DOI] [PubMed] [Google Scholar]
- Piperno G, Luck DJ, 1979. Axonemal adenosine triphosphatases from flagella of Chlamydomonas reinhardtii. Purification of two dyneins. J. Biol. Chem 254 (8), 3084–3090. [PubMed] [Google Scholar]
- Plastino J, Blanchoin L, 2018. Dynamic stability of the actin ecosystem. J. Cell Sci 132 (4). [DOI] [PubMed] [Google Scholar]
- Plattner F, Yarovinsky F, Romero S, Didry D, Carlier M-F, Sher A, Soldati-Favre D, 2008. Toxoplasma profilin is essential for host cell invasion and TLR11-dependent induction of an Interleukin-12 response. Cell Host Microbe 3 (2), 77–87. [DOI] [PubMed] [Google Scholar]
- Plessner M, Knerr J, Grosse R, 2019. Centrosomal actin assembly is required for proper mitotic spindle formation and chromosome congression. iScience 15, 274–281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Polet D, Lambrechts A, Ono K, Mah A, Peelman F, Vandekerckhove J, Baillie DL, Ampe C, Ono S, 2006. Caenorhabditis elegans expresses three functional profilins in a tissue-specific manner. Cell Motil. Cytoskeleton 63 (1), 14–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pollard TD, 1986. Rate constants for the reactions of ATP- and ADP-actin with the ends of actin filaments. J. Cell Biol 103 (6), 2747–2754. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pollard TD, 2007. Regulation of actin filament assembly by Arp2/3 complex and formins. Annu. Rev. Biophys. Biomol. Struct 36, 451–477. [DOI] [PubMed] [Google Scholar]
- Pollard TD, 2016. Actin and actin-binding proteins. Cold Spring Harb. Perspect. Biol 8 (8), a018226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pollard TD, Borisy GG, 2003. Cellular motility driven by assembly and disassembly of actin filaments. Cell 112 (4), 453–465. [DOI] [PubMed] [Google Scholar]
- Pollard TD, O’Shaughnessy B, 2019. Molecular mechanism of cytokinesis. Annu. Rev. Biochem 88, 661–689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pollard TD, Wu J-Q, 2010. Understanding cytokinesis: lessons from fission yeast. Nat. Rev. Mol. Cell Biol 11 (2), 149–155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pollard TD, Blanchoin L, Mullins RD, 2000. Molecular mechanisms controlling actin filament dynamics in nonmuscle cells. Annu Rev. Biophys. Biomol. Struct 29, 545–576. [DOI] [PubMed] [Google Scholar]
- Popov AV, Severin F, Karsenti E, 2002. XMAP215 is required for the microtubule-nucleating activity of centrosomes. Curr. Biol 12 (15), 1326–1330. [DOI] [PubMed] [Google Scholar]
- Posey AE, Ruff KM, Harmon TS, Crick SL, Li A, Diamond MI, Pappu RV, 2018. Profilin reduces aggregation and phase separation of huntingtin N-terminal fragments by preferentially binding to soluble monomers and oligomers. J. Biol. Chem 293 (10), 3734–3746. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prezel E, Elie A, Delaroche J, Stoppin-Mellet V, Bosc C, Serre L, Fourest-Lieuvin A, Andrieux A, Vantard M, Arnal I, 2018. Tau can switch microtubule network organizations: from random networks to dynamic and stable bundles. Mol. Biol. Cell 29 (2), 154–165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prokop A, Beaven R, Qu Y, Sánchez-Soriano N, 2013. Using fly genetics to dissect the cytoskeletal machinery of neurons during axonal growth and maintenance. J. Cell Sci 126 (11), 2331–2341. [DOI] [PubMed] [Google Scholar]
- Pruyne D, Evangelista M, Yang C, Bi E, Zigmond S, Bretscher A, Boone C, 2002. Role of formins in actin assembly: nucleation and barbed-end association. Science 297 (5581), 612–615. [DOI] [PubMed] [Google Scholar]
- Qiao Z, Sun H, Ng JTY, Ma Q, Koh SH, Mu Y, Miao Y, Gao Y-G, 2019. Structural and computational examination of the Arabidopsis profilin–poly-p complex reveals mechanistic details in profilin-regulated actin assembly. J. Biol. Chem 294 (49), 18650–18661. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quinlan ME, Heuser JE, Kerkhoff E, Mullins RD, 2005. Drosophila spire is an actin nucleation factor. Nature 433 (7024), 382–388. [DOI] [PubMed] [Google Scholar]
- Rebowski G, Boczkowska M, Drazic A, Ree R, Goris M, Arnesen T, Dominguez R, 2020. Mechanism of actin N-terminal acetylation. Sci. Adv 6 (15), eaay8793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reeve SP, Bassetto L, Genova GK, Kleyner Y, Leyssen M, Jackson FR, Hassan BA, 2005. The Drosophila fragile X mental retardation protein controls actin dynamics by directly regulating profilin in the brain. Curr. Biol 15 (12), 1156–1163. [DOI] [PubMed] [Google Scholar]
- Reinhard M, Giehl K, Abel K, Haffner C, Jarchau T, Hoppe V, Jockusch BM, Walter U, 1995. The proline-rich focal adhesion and microfilament protein VASP is a ligand for profilins. EMBO J. 14 (8), 1583–1589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ricketts SN, Francis ML, Farhadi L, Rust MJ, Das M, Ross JL, Robertson-Anderson RM, 2019. Varying crosslinking motifs drive the mesoscale mechanics of actin-microtubule composites. Sci. Rep 9 (1), 12831. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rizvi SA, Neidt EM, Cui J, Feiger Z, Skau CT, Gardel ML, Kozmin SA, Kovar DR, 2009. Identification and characterization of a small molecule inhibitor of formin-mediated actin assembly. Chem. Biol 16 (11), 1158–1168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rizwani W, Fasim A, Sharma D, Reddy DJ, Bin Omar NAM, Singh SS, 2014. S137 phosphorylation of profilin 1 is an important signaling event in breast cancer progression. PLoS One 9 (8), e103868. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rodal AA, Manning AL, Goode BL, Drubin DG, 2003. Negative regulation of yeast WASp by two SH3 domain-containing proteins. Curr. Biol 13 (12), 1000–1008. [DOI] [PubMed] [Google Scholar]
- Rodionov VI, Hope AJ, Svitkina TM, Borisy GG, 1998. Functional coordination of microtubule-based and actin-based motility in melanophores. Curr. Biol 8 (3), 165–169. [DOI] [PubMed] [Google Scholar]
- Rodriguez OC, Schaefer AW, Mandato CA, Forscher P, Bement WM, Waterman-Storer CM, 2003. Conserved microtubule-actin interactions in cell movement and morphogenesis. Nat. Cell Biol 5 (7), 599–609. [DOI] [PubMed] [Google Scholar]
- Roeles J, Tsiavaliaris G, 2019. Actin-microtubule interplay coordinates spindle assembly in human oocytes. Nat. Commun 10 (1), 4651. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Romero S, Le Clainche C, Didry D, Egile C, Pantaloni D, Carlier M-F, 2004. Formin is a processive motor that requires profilin to accelerate actin assembly and associated ATP hydrolysis. Cell 119 (3), 419–429. [DOI] [PubMed] [Google Scholar]
- Roostalu J, Surrey T, 2017. Microtubule nucleation: beyond the template. Nat. Rev. Mol. Cell Biol 18 (11), 702–710. [DOI] [PubMed] [Google Scholar]
- Rosenblatt J, Cramer LP, Baum B, McGee KM, 2004. Myosin II-dependent cortical movement is required for centrosome separation and positioning during mitotic spindle assembly. Cell 117 (3), 361–372. [DOI] [PubMed] [Google Scholar]
- Roth LW, Bormann P, Bonnet A, Reinhard E, 1999. Beta-thymosin is required for axonal tract formation in developing zebrafish brain. Development 126 (7), 1365–1374. [DOI] [PubMed] [Google Scholar]
- Roth-Johnson EA, Vizcarra CL, Bois JS, Quinlan ME, 2014. Interaction between microtubules and the Drosophila formin cappuccino and its effect on actin assembly. J. Biol. Chem 289 (7), 4395–4404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rotty JD, Wu C, Haynes EM, Suarez C, Winkelman JD, Johnson HE, Haugh JM, Kovar DR, Bear JE, 2015. Profilin-1 serves as a gatekeeper for actin assembly by Arp2/3-dependent and -independent pathways. Dev. Cell 32 (1), 54–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rouiller I, Xu X-P, Amann KJ, Egile C, Nickell S, Nicastro D, Li R, Pollard TD, Volkmann N, Hanein D, 2008. The structural basis of actin filament branching by the Arp2/3 complex. J. Cell Biol 180 (5), 887–895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roy P, Jacobson K, 2004. Overexpression of profilin reduces the migration of invasive breast cancer cells. Cell Motil. Cytoskeleton 57 (2), 84–95. [DOI] [PubMed] [Google Scholar]
- Rust MJ, Bates M, Zhuang X, 2006. Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM). Nat. Methods 3 (10), 793–796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Safer D, Elzinga M, Nachmias VT, 1991. Thymosin-beta-4 and Fx, an actin-sequestering peptide, are indistinguishable. J. Biol. Chem 266 (7), 4029–4032. [PubMed] [Google Scholar]
- Sagot I, Rodal AA, Moseley J, Goode BL, Pellman D, 2002. An actin nucleation mechanism mediated by Bni1 and profilin. Nat. Cell Biol 4 (8), 626–631. [DOI] [PubMed] [Google Scholar]
- Salmon WC, Adams MC, Waterman-Storer CM, 2002. Dual-wavelength fluorescent speckle microscopy reveals coupling of microtubule and actin movements in migrating cells. J. Cell Biol 158 (1), 31–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sanger JW, 1975. Changing patterns of actin localization during cell division. Proc. Natl. Acad. Sci 72 (5), 1913–1916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Santos A, Van Ree R, 2011. Profilins: mimickers of allergy or relevant allergens. Int. Arch. Allergy Immunol 155 (3), 191–204. [DOI] [PubMed] [Google Scholar]
- Sathish K, Padma B, Munugalavadla V, Bhargavi V, Radhika KVN, Wasia R, Sairam M, Singh SS, 2004. Phosphorylation of profilin regulates its interaction with actin and poly(L)-proline. Cell. Signal 16 (5), 589–596. [DOI] [PubMed] [Google Scholar]
- Schlett K, 2017. More than a mere supply of monomers: G-actin pools regulate actin dynamics in dendritic spines. J. Cell Biol 216 (8), 2255–2257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schlüter K, Jockusch BM, Rothkegel M, 1997. Profilins as regulators of actin dynamics. Biochim. Biophys. Acta 1359 (2), 97–109. [DOI] [PubMed] [Google Scholar]
- Schoppmeyer R, Zhao R, Cheng H, Hamed M, Liu C, Zhou X, Schwarz EC, Zhou Y, Knörck A, Schwär G, Ji S, Liu L, Long J, Helms V, Hoth M, Yu X, Qu B, 2017. Human profilin-1 is a negative regulator of CTL mediated cell-killing and migration. Eur. J. Immunol 47 (9), 1562–1572. [DOI] [PubMed] [Google Scholar]
- Schutt CE, Myslik JC, Rozycki MD, Goonesekere NCW, Lindber U, 1993. The structure of crystalline profilin-β-actin. Nature 365, 810–816. [DOI] [PubMed] [Google Scholar]
- Sellers JR, Shi S, Nishimura Y, Zhang F, Liu R, Takagi Y, Viasnoff V, Bershadsky AD, 2020. The Formin inhibitor, SMIFH2, inhibits members of the myosin superfamily. Biophys. J 118 (3), 125a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sept D, McCammon JA, 2001. Thermodynamics and kinetics of actin filament nucleation. Biophys. J 81 (2), 667–674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Severson AF, Baillie DL, Bowerman B, 2002. A Formin homology protein and a profilin are required for cytokinesis and Arp2/3-independent assembly of cortical microfilaments in C. elegans. Curr. Biol 12 (24), 2066–2075. [DOI] [PubMed] [Google Scholar]
- Shao J, Diamond MI, 2012. Protein Phosphatase-1 dephosphorylates Profilin-1 at ser-137. PLoS One 7 (3), e32802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shao J, Welch WJ, Diprospero NA, Diamond MI, 2008. Phosphorylation of profilin by ROCK1 regulates polyglutamine aggregation. Mol. Cell. Biol 28 (17), 5196–5208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sherer LA, Zweifel ME, Courtemanche N, 2018. Dissection of two parallel pathways for formin-mediated actin filament elongation. J. Biol. Chem 293 (46), 17917–17928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shields AR, Spence AC, Yamashita YM, Davies EL, Fuller MT, 2014. The actin-binding protein profilin is required for germline stem cell maintenance and germ cell enclosure by somatic cyst cells. Development 141 (1), 73–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shirey CM, Scott JL, Stahelin RV, 2017. Notes and tips for improving quality of lipid-protein overlay assays. Anal. Biochem 516, 9–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simons M, Schwartz MA, 2012. Profilin phosphorylation as a VEGFR effector in angiogenesis. Nat. Cell Biol 14 (10), 985–987. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singh SS, Chauhan A, Murakami N, Chauhan VP, 1996. Profilin and gelsolin stimulate phosphatidylinositol 3-kinase activity. Biochem. 35 (51), 16544–16549. [DOI] [PubMed] [Google Scholar]
- Skillman KM, Daher W, Ma CI, Soldati-Favre D, Sibley LD, 2012. Toxoplasma gondii profilin acts primarily to sequester G-actin while formins efficiently nucleate actin filament formation in vitro. Biochem. 51 (12), 2486–2495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Skruber K, Read T-A, Vitriol EA, 2018. Reconsidering an active role for G-actin in cytoskeletal regulation. J. Cell Sci 131 (1). [DOI] [PMC free article] [PubMed] [Google Scholar]
- Skruber K, Warp PV, Shklyarov R, Thomas JD, Swanson MS, Henty-Ridilla JL, Read T-A, Vitriol EA, 2020. Arp2/3 and Mena/VASP require Profilin 1 for actin network assembly at the leading edge, Curr. Biol 30, 1–14.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Slater PG, Hayrapetian L, Lowery LA, 2017. Xenopus laevis as a model system to study cytoskeletal dynamics during axon pathfinding. Genesis 55 (1–2), e22994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Small JV, 1988. The actin cytoskeleton. Electron Microsc. Rev 1 (1), 155–174. [DOI] [PubMed] [Google Scholar]
- Small JV, Celis JE, 1978. Filament arrangements in negatively stained cultured cells: the organization of actin. Cytobiologie 16 (2), 308–325. [PubMed] [Google Scholar]
- Smith TE, Hong W, Zachariah MM, Harper MK, Matainaho TK, Van Wagoner RM, Ireland CM, Vershinin M, 2013. Single-molecule inhibition of human kinesin by adociasulfate-13 and −14 from the sponge Cladocroce aculeata. Proc. Natl. Acad. Sci 110 (47), 18880–18885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Söderberg E, Hessle V, von Euler A, Visa N, 2012. Profilin is associated with transcriptionally active genes. Nucleus 3 (3), 290–299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sohn RH, Chen J, Koblan KS, Bray PF, Goldschmidt-Clermont PJ, 1995. Localization of a binding site for phosphatidylinositol 4,5-bisphosphate on human profilin. J. Biol. Chem 270 (36), 21114–21120. [DOI] [PubMed] [Google Scholar]
- Staiger CJ, Sheahan MB, Khurana P, Wang X, McCurdy DW, Blanchoin L, 2009. Actin filament dynamics are dominated by rapid growth and severing activity in the Arabidopsis cortical array. J. Cell Biol 184 (2), 269–280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stehn JR, Haass NK, Bonello T, Desouza M, Kottyan G, Treutlein H, Zeng J, Nascimento PRBB, Sequeira VB, Butler TL, Allanson M, Fath T, Hill TA, McCluskey A, Schevzov G, Palmer SJ, Hardeman EC, Winlaw D, Reeve VE, Dixon I, Weninger W, Cripe TP, Gunning PW, 2013. A novel class of anticancer compounds targets the actin cytoskeleton in tumor cells. Cancer Res. 73 (16), 5169–5182. [DOI] [PubMed] [Google Scholar]
- Sturgill EG, Norris SR, Guo Y, Ohi R, 2016. Kinesin-5 inhibitor resistance is driven by kinesin-12. J. Cell Biol 213 (2), 213–227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stüven T, Hartmann E, Görlich D, 2003. Exportin 6: a novel nuclear export receptor that is specific for profilin-actin complexes. EMBO J. 22 (21), 5928–5940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suarez C, Kovar DR, 2016. Internetwork competition for monomers governs actin cytoskeleton organization. Nat. Rev. Mol. Cell Biol 17 (12), 799–810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suarez C, Carroll RT, Burke TA, Christensen JR, Bestul AJ, Sees JA, James ML, Sirotkin V, Kovar DR, 2015. Profilin regulates F-actin network homeostasis by favoring formin over Arp2/3 complex. Dev. Cell 32 (1), 43–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suetsugu S, Miki H, Takenawa T, 1998. The essential role of profilin in the assembly of actin for microspike formation. EMBO J. 17 (22), 6516–6526. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun H, Qiao Z, Chua KP, Tursic A, Liu X, Gao Y-G, Mu Y, Hou X, Miao Y, 2018. Profilin negatively regulates Formin-mediated actin assembly to modulate PAMP-triggered plant immunity. Curr. Biol 28 (12), 1882–1895.e7. [DOI] [PubMed] [Google Scholar]
- Svitkina TM, 2018. Ultrastructure of the actin cytoskeleton. Curr. Opin. Cell Biol 54, 1–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Svitkina TM, Borisy GG, 1999. Arp2/3 complex and actin depolymerizing factor/cofilin in dendritic organization and treadmilling of actin filament array in lamellipodia. J. Cell Biol 145 (5), 1009–1026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Svitkina TM, Verkhovsky AB, McQuade KM, Borisy GG, 1997. Analysis of the actin-myosin II system in fish epidermal keratocytes: mechanism of cell body translocation. J. Cell Biol 139 (2), 397–415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Svitkina TM, Bulanova EA, Chaga OY, Vignjevic DM, Kojima S, Vasiliev JM, Borisy GG, 2003. Mechanism of filopodia initiation by reorganization of a dendritic network. J. Cell Biol 160 (3), 409–421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Symons MH, Mitchison TJ, 1991. Control of actin polymerization in live and permeabilized fibroblasts. J. Cell Biol 114 (3), 503–513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Szikora S, Földi I, Tóth K, Migh E, Vig A, Bugyi B, Maléth J, Hegyi P, Kaltenecker P, Sanchez-Soriano N, Mihály J, 2017. The formin DAAM is required for coordination of the actin and microtubule cytoskeleton in axonal growth cones. J. Cell Sci 130, 2506–2519. [DOI] [PubMed] [Google Scholar]
- Tai AW, Chuang JZ, Bode C, Wolfrum U, Sung CH, 1999. Rhodopsin’s carboxy-terminal cytoplasmic tail acts as a membrane receptor for cytoplasmic dynein by binding to the dynein light chain Tctex-1. Cell 97 (7), 877–887. [DOI] [PubMed] [Google Scholar]
- Tang Y-N, Ding W-Q, Guo X-J, Yuan X-W, Wang D-M, Song J-G, 2015. Epigenetic regulation of Smad2 and Smad3 by profilin-2 promotes lung cancer growth and metastasis. Nat. Commun 6, 8230. [DOI] [PubMed] [Google Scholar]
- Tas RP, Chazeau A, Cloin BMC, Lambers MLA, Hoogenraad CC, Kapitein LC, 2017. Differentiation between oppositely oriented microtubules controls polarized neuronal transport. Neuron 96 (6), 1264–1271.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thawani A, Kadzik RS, Petry S, 2018. XMAP215 is a microtubule nucleation factor that functions synergistically with the γ-tubulin ring complex. Nat. Cell Biol 20 (5), 575–585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Theriot JA, Mitchison TJ, Tilney LG, Portnoy DA, 1992. The rate of actin-based motility of intracellular listeria monocytogenes equals the rate of actin polymerization. Nature 357 (6375), 257–260. [DOI] [PubMed] [Google Scholar]
- Theriot JA, Rosenblatt J, Portnoy DA, Goldschmidt-Clermont PJ, Mitchison TJ, 1994. Involvement of profilin in the actin-based motility of L. monocytogenes in cells and in cell-free extracts. Cell 76 (3), 505–517. [DOI] [PubMed] [Google Scholar]
- Théry M, Racine V, Pépin A, Piel M, Chen Y, Sibarita J-B, Bornens M, 2005. The extracellular matrix guides the orientation of the cell division axis. Nat. Cell Biol 7 (10), 947–953. [DOI] [PubMed] [Google Scholar]
- Thorn KS, Christensen HE, Shigeta R, Huddler D, Shalaby L, Lindberg U, Chua NH, Schutt CE, 1997. The crystal structure of a major allergen from plants. Structure 5 (1), 19–32. [DOI] [PubMed] [Google Scholar]
- Tilney LG, 1976. The polymerization of actin. Aggregates of nonfilamentous actin and its associated proteins: a storage form of actin. J. Cell Biol 69 (1), 73–89. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tilney LG, DeRosier DJ, Weber A, Tilney MS, 1992. How Listeria exploits host cell actin to form its own cytoskeleton. Nucleation, actin filament polarity, filament assembly, and evidence for a pointed end capper. J. Cell Biol 118 (1), 83–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toyoshima F, Nishida E, 2007. Spindle orientation in animal cell mitosis: roles of integrin in the control of spindle axis. J. Cell. Physiol 213 (2), 407–411. [DOI] [PubMed] [Google Scholar]
- Ubersax JA, Woodbury EL, Quang PN, Paraz M, Blethrow JD, Shah K, Shokat KM, Morgan DO, 2003. Targets of the cyclin-dependent kinase Cdk1. Nature 425 (6960), 859–864. [DOI] [PubMed] [Google Scholar]
- Vargas P, Barbier L, Sáez PJ, Piel M, 2017. Mechanisms for fast cell migration in complex environments. Curr. Opin. Cell Biol 48, 72–78. [DOI] [PubMed] [Google Scholar]
- Verheyen EM, Cooley L, 1994. Profilin mutations disrupt multiple actin-dependent processes during Drosophila development. Development 120 (4), 717–728. [DOI] [PubMed] [Google Scholar]
- Vidali L, Augustine RC, Kleinman KP, Bezanilla M, 2007. Profilin is essential for tip growth in the moss Physcomitrella patens. Plant Cell 19 (11), 3705–3722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vignaud T, Copos C, Leterrier C, Tseng Q, Blanchoin L, Mogilner A, Théry M, Kurzawa L, 2020. Stress fibers are embedded in a contractile cortical network. bioRxiv 10.1101/2020.02.11.944579, Preprint. [DOI] [PMC free article] [PubMed]
- Vinson V, Archer S, Lattman E, Pollard T, Torchia D, 1993. Three-dimensional solution structure of Acanthamoeba profilin-I. J. Cell Biol 122 (6), 1277–1283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vinson VK, De La Cruz EM, Higgs HN, Pollard TD, 1998. Interactions of Acanthamoeba profilin with actin and nucleotides bound to actin. Biochem. 37 (31), 10871–10880. [DOI] [PubMed] [Google Scholar]
- Vitriol EA, McMillen LM, Kapustina M, Gomez SM, Vavylonis D, Zheng JQ, 2015. Two functionally distinct sources of actin monomers supply the leading edge of lamellipodia. Cell Rep. 11 (3), 433–445. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Voter WA, Erickson HP, 1984. The kinetics of microtubule assembly. Evidence for a two-stage nucleation mechanism. J. Biol. Chem 259 (16), 10430–10438. [PubMed] [Google Scholar]
- Wade RH, 2007. Microtubules: an overview. Meth. Mol. Med 137, 1–16. [DOI] [PubMed] [Google Scholar]
- Walter LM, Franz P, Lindner R, Tsiavaliaris G, Hensel N, Claus P, 2020. Profilin2a-phosphorylation as a regulatory mechanism for actin dynamics. FASEB J. 34 (2), 2147–2160. [DOI] [PubMed] [Google Scholar]
- Wang YL, 1984. Reorganization of actin filament bundles in living fibroblasts. J. Cell Biol 99 (4), 1478–1485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wase N, Tu B, Rasineni GK, Cerny R, Grove R, Adamec J, Black PN, DiRusso CC, 2019. Remodeling of Chlamydomonas metabolism using synthetic inducers results in lipid storage during growth. Plant Physiol. 181 (3), 1029–1049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Watanabe N, Madaule P, Reid T, Ishizaki T, Watanabe G, Kakizuka A, Saito Y, Nakao K, Jockusch BM, Narumiya S, 1997. p140mDia, a mammalian homolog of Drosophila diaphanous, is a target protein for rho small GTPase and is a ligand for profilin. EMBO J. 16 (11), 3044–3056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Watanabe S, Ando Y, Yasuda S, Hosoya H, Watanabe N, Ishizaki T, Narumiya S, 2008. mDia2 induces the actin scaffold for the contractile ring and stabilizes its position during cytokinesis in NIH 3T3 cells. Mol. Biol. Cell 19 (5), 2328–2338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Waterman-Storer CM, Salmon ED, 1998. How microtubules get fluorescent speckles. Biophys. J 75 (4), 2059–2069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weaver BA, 2014. How Taxol/paclitaxel kills cancer cells. Mol. Biol. Cell 25 (18),2677–2681. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wickramarachchi DC, Theofilopoulos AN, Kono DH, 2010. Immune pathology associated with altered actin cytoskeleton regulation. J. Autoimmun 43 (1), 64–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wieczorek M, Bechstedt S, Chaaban S, Brouhard GJ, 2015. Microtubule-associated proteins control the kinetics of microtubule nucleation. Nat. Cell Biol 17 (7), 907–916. [DOI] [PubMed] [Google Scholar]
- Wiesner S, Helfer E, Didry D, Ducouret G, Lafuma F, Carlier M-F, Pantaloni D, 2003. A biomimetic motility assay provides insight into the mechanism of actin-based motility. J. Cell Biol 160 (3), 387–398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilson NF, Foglesong MJ, Snell WJ, 1997. The Chlamydomonas mating type plus fertilization tubule, a prototypic cell fusion organelle: isolation, characterization, and in vitro adhesion to mating type minus gametes. J. Cell Biol 137 (7), 1537–1553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Witke W, 2004. The role of profilin complexes in cell motility and other cellular processes. Trends Cell Biol. 14, 461–469. [DOI] [PubMed] [Google Scholar]
- Witke W, Podtelejnikov AV, Di Nardo A, Sutherland JD, Gurniak CB, Dotti C, Mann M, 1998. In mouse brain profilin I and profilin II associate with regulators of the endocytic pathway and actin assembly. EMBO J. 17 (4), 967–976. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Witke W, Sutherland JD, Sharpe A, Arai M, Kwiatkowski DJ, 2001. Profilin I is essential for cell survival and cell division in early mouse development. Proc. Natl. Acad. Sci 98 (7), 3832–3836. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wittenmayer N, Rothkegel M, Jockusch BM, Schlüter K, 2000. Functional characterization of green fluorescent protein-profilin fusion proteins: GFP-profilin fusion proteins. Eur. J. Biochem 267 (16), 5247–5256. [DOI] [PubMed] [Google Scholar]
- Wittenmayer N, Jandrig B, Rothkegel M, Schlüter K, Arnold W, Haensch W, Scherneck S, Jockusch BM, 2004. Tumor suppressor activity of profilin requires a functional actin binding site. Mol. Biol. Cell 15 (4), 1600–1608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wittmann T, Waterman-Storer CM, 2001. Cell motility: Can Rho GTPases and microtubules point the way?. J. Cell Sci 114 (21), 3795–3803. [DOI] [PubMed] [Google Scholar]
- Wolven AK, Belmont LD, Mahoney NM, Almo SC, Drubin DG, 2000. In vivo importance of actin nucleotide exchange catalyzed by profilin. J. Cell Biol 150 (4), 895–904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu C-H, Fallini C, Ticozzi N, Keagle PJ, Sapp PC, Piotrowska K, Lowe P, Koppers M, McKenna-Yasek D, Baron DM, Kost JE, Gonzalez-Perez P, Fox AD, Adams J, Taroni F, Tiloca C, Leclerc AL, Chafe SC, Mangroo D, … Landers JE, 2012. Mutations in the profilin 1 gene cause familial amyotrophic lateral sclerosis. Nature 488 (7412), 499–503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamashiro S, Yamakita Y, Hosoya H, Matsumura F, 1991. Phosphorylation of non-muscle caldesmon by p34cdc2 kinase during mitosis. Nature 349 (6305), 169–172. [DOI] [PubMed] [Google Scholar]
- Yang D, Wang Y, Jiang M, Deng X, Pei Z, Li F, Xia K, Zhu L, Yang T, Chen M, 2017. Downregulation of Profilin-1 expression attenuates cardiomyocytes hypertrophy and apoptosis induced by advanced glycation end products in H9c2 cells. Biomed. Res. Int 2017, 9716087. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yao W, Ji S, Qin Y, Yang J, Xu J, Zhang B, Xu W, Liu J, Shi S, Liu L, Liu C, Long J, Ni Q, Li M, Yu X, 2014. Profilin-1 suppresses tumorigenicity in pancreatic cancer through regulation of the SIRT3-HIF1α axis. Mol. Cancer 13, 187. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yarmola EG, Parikh S, Bubb MR, 2001. Formation and implications of a ternary complex of profilin, thymosin beta 4, and actin. J. Biol. Chem 276 (49), 45555–45563. [DOI] [PubMed] [Google Scholar]
- Yarovinsky F, Zhang D, Andersen JF, Bannenberg GL, Serhan CN, Hayden MS, Hieny S, Sutterwala FS, Flavell RA, Ghosh S, Sher A, 2005. TLR11 activation of dendritic cells by a protozoan profilin-like protein. Science 308 (5728), 1626–1629. [DOI] [PubMed] [Google Scholar]
- Yuan X, Song M, Devine P, Bruneau BG, Scott IC, Wilson MD, 2018. Heart enhancers with deeply conserved regulatory activity are established early in zebrafish development. Nat. Commun 9 (1), 4977. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zaremba-Niedzwiedzka K, Caceres EF, Saw JH, Bäckström D, Juzokaite L, Vancaester E, Seitz KW, Anantharaman K, Starnawski P, Kjeldsen KU, Stott MB, Nunoura T, Banfield JF, Schramm A, Baker BJ, Spang A, Ettema TJG, 2017. Asgard archaea illuminate the origin of eukaryotic cellular complexity. Nature 541 (7637), 353–358. [DOI] [PubMed] [Google Scholar]
- Zhang R, Alushin GM, Brown A, Nogales E, 2015. Mechanistic origin of microtubule dynamic instability and its modulation by EB proteins. Cell 162 (4), 849–859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang H, Yang W, Yan J, Zhou K, Wan B, Shi P, Chen Y, He S, Li D, 2018. Loss of profilin 2 contributes to enhanced epithelial-mesenchymal transition and metastasis of colorectal cancer. Int. J. Oncol 53 (3), 1118–1128. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou K, Chen J, Wu J, Xu Y, Wu Q, Yue J, Song Y, Li S, Zhou P, Tu W, Yang G, Jiang S, 2019. Profilin 2 promotes proliferation and metastasis of head and neck cancer cells by regulating PI3K/AKT/β-catenin signaling pathway. Oncol. Res 27 (9), 1079–1088. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zoidakis J, Makridakis M, Zerefos PG, Bitsika V, Esteban S, Frantzi M, Stravodimos K, Anagnou NP, Roubelakis MG, Sanchez-Carbayo M, Vlahou A, 2012. Profilin 1 is a potential biomarker for bladder cancer aggressiveness. Mol. Cell. Proteomics 11 (4), M111.009449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zou L, Jaramillo M, Whaley D, Wells A, Panchapakesa V, Das T, Roy P, 2007. Profilin-1 is a negative regulator of mammary carcinoma aggressiveness. Brit. J. Cancer 97 (10), 1361–1371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zou L, Ding Z, Roy P, 2010. Profilin-1 overexpression inhibits proliferation of MDA-MB-231 breast cancer cells partly through p27kip1 upregulation. J. Cell. Physiol 223 (3), 623–629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zweifel ME, Courtemanche N, 2020. Competition for delivery of profilin-actin to barbed ends limits the rate of formin-mediated actin filament elongation. J. Biol. Chem 295, 4513–4525. [DOI] [PMC free article] [PubMed] [Google Scholar]