Home Common mechanisms of catalysis in small and heterotrimeric GTPases and their respective GAPs
Article Open Access

Common mechanisms of catalysis in small and heterotrimeric GTPases and their respective GAPs

  • Klaus Gerwert EMAIL logo , Daniel Mann and Carsten Kötting EMAIL logo
Published/Copyright: February 23, 2017

Abstract

GTPases are central switches in cells. Their dysfunctions are involved in severe diseases. The small GTPase Ras regulates cell growth, differentiation and apoptosis by transmitting external signals to the nucleus. In one group of oncogenic mutations, the ‘switch-off’ reaction is inhibited, leading to persistent activation of the signaling pathway. The switch reaction is regulated by GTPase-activating proteins (GAPs), which catalyze GTP hydrolysis in Ras, and by guanine nucleotide exchange factors, which catalyze the exchange of GDP for GTP. Heterotrimeric G-proteins are activated by G-protein coupled receptors and are inactivated by GTP hydrolysis in the Gα subunit. Their GAPs are called regulators of G-protein signaling. In the same way that Ras serves as a prototype for small GTPases, Gαi1 is the most well-studied Gα subunit. By utilizing X-ray structural models, time-resolved infrared-difference spectroscopy, and biomolecular simulations, we elucidated the detailed molecular reaction mechanism of the GTP hydrolysis in Ras and Gαi1. In both proteins, the charge distribution of GTP is driven towards the transition state, and an arginine is precisely positioned to facilitate nucleophilic attack of water. In addition to these mechanistic details of GTP hydrolysis, Ras dimerization as an emerging factor in signal transduction is discussed in this review.

Introduction: small and heterotrimeric GTPases

The overall structure of and common motifs present in small and heterotrimeric GTPases have been previously discussed in several excellent reviews (Hamm, 1998; Vetter and Wittinghofer, 2001; Wennerberg et al., 2005; Wittinghofer and Vetter, 2011; Ligeti et al., 2012; Cherfils and Zeghouf, 2013; Wittinghofer, 2014a,b; Carvalho et al., 2015; Lu et al., 2016; Mishra and Lambright, 2016; Sprang, 2016; Syrovatkina et al., 2016). Here, we will focus mainly on the results obtained from the orchestration of time-resolved Fourier transform infrared (FTIR) difference-spectroscopy experiments and biomolecular simulations. These studies were performed within the collaborative research center 642 of the Deutsche Forschungsgemeinschaft (DFG), and the experimental methods were previously described in detail in Kötting and Gerwert (2015).

GTPases are molecular switches that regulate a plethora of cellular events (Brunsveld et al., 2006; Cherfils and Zeghouf, 2013). The signaling state is encoded into surface alterations of the GTPase that are determined by the presence of the third phosphate group of GTP (Wittinghofer and Vetter, 2011). GTP hydrolysis, which switches the signal off, has been the focus of extensive research spanning dozens of years. When this reaction occurs in water, it is very slow, with a half-life of about 60 days at 37°C (Kötting and Gerwert, 2004). However, a typical small GTPase, such as Ras, catalyzes GTP hydrolysis with an efficiency that exceeds this by five orders of magnitude, to t½=25 min at 37°C (Tucker et al., 1986; Neal et al., 1990). GTPase-activating proteins (GAPs) can further accelerate hydrolysis by forming a protein-protein complex with the GTPase; in the case of GAP-Ras, the hydrolysis is catalyzed by another five orders of magnitude, down to t½=36 ms at 25°C (Gideon et al., 1992).

In contrast, the Gα subunits of heterotrimeric GTPases can hydrolyze GTP in seconds, which is much faster than the reaction mediated by small GTPases (Linder et al., 1990). Many Gα subunits are also catalyzed by GAPs, known as regulators of G-protein signaling (RGS) proteins, and these promote reaction rates similar to those of the Ras-GAP complex. Importantly, mutations that interfere with GTP hydrolysis are frequent in various tumors (Pylayeva-Gupta et al., 2011; O’Hayre et al., 2013) and fibrous dysplasia (Dumitrescu and Collins, 2008). Other diseases result from modifications of catalytic residues within the binding pocket, such as ADP-ribosylation of Gα subunits by cholera toxin (Northup et al., 1980).

All GTPases share a G-domain with five α-helices and a six-stranded β-sheet. They form the two large families described above, the heterotrimeric GTPases (Sprang, 2016) and the small GTPases (Mishra and Lambright, 2016), as well as the dynamin superfamily (Daumke and Praefcke, 2016) and translational GTPases (Maracci and Rodnina, 2016). The small GTPases are grouped into five subfamilies: Ras GTPases are mainly involved in cell growth (Cox and Der, 2010), Rho GTPases function in the regulation of the cytoskeleton (Sit and Manser, 2011), Rab (Stenmark, 2009) and Arf (Khan and Ménétrey, 2013) regulate vesicular transport, and Ran (Jamali et al., 2011) controls nuclear transport. In contrast, heterotrimeric G-proteins are involved in mediating vision, smell, or regulation of the cAMP second messenger system, that modulates among others the blood pressure (Walland, 1975). To date, several different α-, β- and γ-subunits, that can form a heterotrimer, are known, most prominently Gαt, which mediates vision, and Gαi and Gαs, which inhibit and stimulate adenylate cyclase, respectively. The Gα subunit, the actual GTPase, consists of a so-called Ras-like domain with high structural similarity to Ras proteins, and in addition a helical all-α domain (Figure 1A). Several conserved active site elements catalyze GTP hydrolysis in small and heterotrimeric GTPases. A conserved Gln/Asn coordinates the attacking water molecule and the charge distribution of the substrate is shifted, mainly by a bound Mg2+-ion, a conserved lysine (GxxxGKS/T motif) and an arginine ‘finger’ (Figure 1B). In this article, several findings on how this active site elements catalyze GTP hydrolysis are summarized.

Figure 1: Structural similarities between Ras and Gαi1.(A) The Gα subunit of heterotrimeric G-proteins consists of a Ras-like domain (orange) and in addition a helical all-Alpha domain (yellow). Ras is shown in blue. (B) The active sites of the Ras·GAP and Gαi1·RGS protein complexes are shown. The active sites share several highly conserved structural elements, most prominently a conserved lysine in the P-loop and the arginine ‘finger’, which is a residue introduced by the GAP in case of Ras but an intrinsic residue from the all-Alpha domain in Gα i1, allosterically positioned by RGS.
Figure 1:

Structural similarities between Ras and Gαi1.

(A) The Gα subunit of heterotrimeric G-proteins consists of a Ras-like domain (orange) and in addition a helical all-Alpha domain (yellow). Ras is shown in blue. (B) The active sites of the Ras·GAP and Gαi1·RGS protein complexes are shown. The active sites share several highly conserved structural elements, most prominently a conserved lysine in the P-loop and the arginine ‘finger’, which is a residue introduced by the GAP in case of Ras but an intrinsic residue from the all-Alpha domain in Gα i1, allosterically positioned by RGS.

GTP hydrolysis in Ras

GTP hydrolysis alone is a slow process with a high-energy transition state. The attacking water molecule must cross the negatively charged oxygen atoms for an in-line nucleophilic attack at γ-GTP. The GTPase active site contains various specific amino acids that function to compensate charges on the GTP molecule, coordinate the Mg2+ atom between β-GTP and γ-GTP, and position the attacking water molecule. Two conserved residues in particular, the Lys16 in the P-loop and the arginine ‘finger’, that is supplied by the GAP, alter the charge distribution of β-GTP and γ-GTP (Li and Zhang, 2004).

Ras is involved in many signal transduction processes (Cox and Der, 2010; Wittinghofer and Vetter, 2011) and can be activated by its guanine nucleotide exchange factor (GEF) son of sevenless (SOS), which catalyzes the exchange of GDP for GTP. SOS reduces the affinity of Ras for the nucleotide and the bound Mg2+ ion (Gureasko et al., 2010), allowing the exchange of GDP (off) for GTP (on) and thus, turning the switch on (Boriack-Sjodin et al., 1998). In the GTP bound ‘on’ state, Ras interacts with downstream effector proteins, transducing the signal mostly via the MAPK pathway to the nucleus. The switch-off is mediated by hydrolysis of GTP to GDP, and this is accelerated and thereby regulated by GAPs.

The detailed reaction mechanisms for Ras and the Ras-GAP complex have been elucidated by time-resolved-FTIR difference-spectroscopy (Cepus et al., 1998; Kötting et al., 2008). Integration of results from X-ray crystallography structures and FTIR spectroscopy experiments (Pai, 1990; Scheffzek et al., 1997) provides not only a static picture of the ground state, but also a ‘movie’ of the reaction mechanism at subatomic detail (available at www.bph.rub.de/movies). Specifically, during the hydrolysis reaction, three apparent rate constants were observed in the Ras-GAP complex by time-resolved FTIR difference spectroscopy, as summarized in Figure 2. To reveal the reactions of individual protein and substrate groups in the catalytic center, difference spectra between the initial-state GTP and the reaction intermediates were calculated. In this way, the large background absorbance of all non-contributing protein groups and the solvent is subtracted.

Figure 2: The Ras·GAP GTPase reaction.(A) The reaction mechanism can be described by three apparent rate constants; the shift to the ‘on’ conformation (k1), the movement of the Arg-finger with immediate hydrolysis of GTP (k2), and the release of Pi into the bulk solvent (k3). Note that the given rate constants were measured at 260 K. (B) With k1, the signal transducing ‘on’ state is formed by the switch-I movement, closing the binding site. Only with the last apparent rate, k3, does the protein return to the ‘off’ state. (C) The Arg-finger moves into the binding pocket with k2. With the same rate, k2, bond breakage is observed, and protein-bound H2PO4−is formed. In the rate-limiting step, k3, H2PO4− is released from the protein into the bulk solvent.
Figure 2:

The Ras·GAP GTPase reaction.

(A) The reaction mechanism can be described by three apparent rate constants; the shift to the ‘on’ conformation (k1), the movement of the Arg-finger with immediate hydrolysis of GTP (k2), and the release of Pi into the bulk solvent (k3). Note that the given rate constants were measured at 260 K. (B) With k1, the signal transducing ‘on’ state is formed by the switch-I movement, closing the binding site. Only with the last apparent rate, k3, does the protein return to the ‘off’ state. (C) The Arg-finger moves into the binding pocket with k2. With the same rate, k2, bond breakage is observed, and protein-bound H2PO4is formed. In the rate-limiting step, k3, H2PO4 is released from the protein into the bulk solvent.

In FTIR experiments, to trigger the reaction in situ, caged GTP (cgGTP) is bound to Ras and then uncaged to produce GTP by a short laser flash (Cepus et al., 1998). Thereby all Ras proteins start the GTPase reaction in a highly synchronized manner with their natural substrate GTP (Kötting and Gerwert, 2015). In the first step (with apparent rate constant k1), the absorbance change of the switch-I marker band indicates the ‘off’ to ‘on’ conformational change after uncaging Ras-cgGTP to Ras GTP. In the second step (with apparent rate constant k2), the arginine finger moves into the catalytic site to complete the formation of the active site for catalysis of GTP hydrolysis (Kötting et al., 2008). Concurrently, cleavage of the γ-phosphate bond occurs, and protein-bound Pi is formed as H2PO4 (Figure 2C) (Kötting et al., 2006; Xia et al., 2012). In the last step (with apparent rate constant k3), the protein-bound phosphate is released into the bulk solvent, the switch-I returns to the ‘off’ state, and the arginine finger leaves the binding pocket again. Thus, time-resolved FTIR spectroscopy demonstrated, under near-physiological conditions, that in contrast to the transition state analog, the arginine finger is not located within the GTP binding pocket in the ground state (Scheffzek et al., 1997). This confirmed the model from X-ray structural models of the small GTPase Rho in complex with RhoGAP, where the transition state analog showed the GAP arginine interacting with the nucleotide, whereas in the ground state, the arginine was not in the active site (Rittinger et al., 1997a,b). The movement of the arginine finger into the catalytic site could be observed directly in k2, immediately followed by bond cleavage. Further, Ras returned to the ‘off’ state after Pi was released into the bulk solvent, making Pi release the rate-limiting step in Ras-catalyzed GTP hydrolysis.

Structural details and charge distribution of Ras-bound GTP were further revealed by integration of the experimental data with biomolecular simulations (Rudack et al., 2012). Specifically, the detailed molecular information encoded in the FTIR difference spectra of GTP can be decoded using quantum mechanics/molecular mechanics (QM/MM) simulations. As validation of the simulation accuracy, the agreement between the experimental and simulated infrared (IR)-GTP spectra at the catalytic center was used as very sensitive measure. Based on their excellent agreement, further detailed conclusions were possible. These provide profound and detailed insights, which are summarized in Figure 3. Here, we show the nucleotide GTP in water, GTP bound to Ras, and GTP bound to the Ras-GAP complex. Protein binding induces catalytically important alterations in GTP, and both the charge distribution and geometry are shifted towards the transition state. That is, the charge of the γ-phosphorous atom becomes more positive, facilitating the nucleophilic attack of the water molecule. Additionally, the distance between the β- and γ-phosphorus atoms increases. Most remarkably, a conformational change in the phosphate groups, from a staggered to an eclipsed conformation, induces strain in the GTP molecule. These alterations in the reactant GTP also drive it energetically closer to the transition state and, in turn, lower the activation barrier for GTP hydrolysis (Figure 4) (Kötting and Gerwert, 2004; Kötting et al., 2008). The high intrinsic binding energy of GTP overcompensates for the energy-costly process of inducing these changes (Jencks, 2006).

Figure 3: Comparison of the detailed structures and charge distributions of GTP free in water, bound to Ras, and in the Ras-GAP complex, as obtained by an integration of experimental FTIR spectroscopy and biomolecular simulations (Rudack et al., 2012).
Figure 3:

Comparison of the detailed structures and charge distributions of GTP free in water, bound to Ras, and in the Ras-GAP complex, as obtained by an integration of experimental FTIR spectroscopy and biomolecular simulations (Rudack et al., 2012).

Figure 4: The free energy during GTP hydrolysis within the Ras-GAP complex (Kötting et al., 2008).The structural changes and charge distributions of GTP in water, bound to Ras, and in the Ras-GAP complex are shown in Figure 3. These shift GTP in Ras and Ras-GAP closer to the transition state for bond cleavage, as indicated here. Note that the dotted lines for the smaller systems indicate their positions relative to the first transition state only, neglecting the energy for complex formation. The subsequent reaction profile is only valid for the Ras·GAP system. Formally, this situation corresponds to a preferential binding of the transition state over the ground state within the protein (Roston and Cui, 2016).
Figure 4:

The free energy during GTP hydrolysis within the Ras-GAP complex (Kötting et al., 2008).

The structural changes and charge distributions of GTP in water, bound to Ras, and in the Ras-GAP complex are shown in Figure 3. These shift GTP in Ras and Ras-GAP closer to the transition state for bond cleavage, as indicated here. Note that the dotted lines for the smaller systems indicate their positions relative to the first transition state only, neglecting the energy for complex formation. The subsequent reaction profile is only valid for the Ras·GAP system. Formally, this situation corresponds to a preferential binding of the transition state over the ground state within the protein (Roston and Cui, 2016).

With this elucidation of the catalytic details for the wild-type protein, it will next be important to investigate the catalytic mechanism in GTPases containing oncogenic mutations and, specifically, how these alterations slow the reaction. In this context, the role of the nucleophilic attacking water molecules in particular will be elucidated in detail.

GTP hydrolysis in heterotrimeric GTPases

Heterotrimeric G-proteins are involved in numerous physiological processes, such as vision, scent, and blood pressure regulation (Malbon, 2005). They consist of an α-, β- and γ-subunit. The Gα-subunit contains the G-domain, which is conserved among GTPases, as well as a helical all-α domain. The latter also contributes the arginine finger, which in case of Ras, is supplied by the GAP. During the off-switch reaction, Gα is not bound to the βγ subunits, and thus GTP hydrolysis can be investigated with Gα alone (Coleman et al., 1994). The GEFs associated with heterotrimeric G-proteins are usually G-protein coupled receptors (GPCRs), which catalyze the exchange from GDP to GTP at the Gα subunit and are currently the focus of investigations by numerous research groups worldwide. A seminal advancement in this field was the elucidation of the β-adrenergic receptor in complex with the G protein (Rasmussen et al., 2011). GPCR signaling pathways are important targets in pharmacology. However, not only dysfunctions in GPCRs, but also in heterotrimeric G-proteins, are responsible for a number of serious diseases (O’Hayre et al., 2013).

In Gα proteins, the nucleotide exchange is much slower than the hydrolysis reaction, and thus, multiple turnover measurements consisting of numerous nucleotide exchanges with subsequent hydrolysis almost exclusively probe the nucleotide exchange. This prevents a detailed study of the hydrolysis reaction by multiple turnover assays; for example, the influence of site-directed mutants on the hydrolysis reaction cannot be studied. Consequently, time-resolved FTIR measurements are especially valuable, as they provide unambiguous single turnover reaction rates for the hydrolysis reaction.

We investigated the kinetics of nucleotide exchange and of hydrolysis in Gαi1 separately, using a fluorescence assay for the exchange reaction and time-resolved FTIR difference spectroscopy for the hydrolysis reaction (Schröter et al., 2015). In Figure 5, we show that a mutation of the intrinsic arginine finger, R178S, does not alter the nucleotide exchange rate, but rather slows hydrolysis by two orders of magnitude (O’Hayre et al., 2013). This could not be deduced from multiple turnover measurements where the reaction is barely slowed. Conversely, the mutant D272N lacks an interaction between the G-domain and the base of the nucleotide, which accelerates the nucleotide exchange kinetics, while the hydrolysis kinetics remain unaltered, as demonstrated by single-turnover FTIR measurements. By integrating time-resolved FTIR difference spectroscopy and molecular dynamics (MD) simulations, new details of the nucleotide exchange reaction can be elucidated. For example, we see that residue D229 positions R242 of the Ras-like G-domain, enabling a stable hydrogen bond to Q147 in the all-α domain. Further, we observe that D231 forms a hydrogen bond to R144. Thereby a detailed picture of the inter-domain coordination, that regulates the nucleotide exchange, emerges (Schröter et al., 2015).

Figure 5: Multiple turnover kinetics of GDP to GTP exchange and GTP hydrolysis under steady-state conditions in Gαi1 (left), and the underlying kinetics of single turnover nucleotide exchange (middle) and single turnover GTP hydrolysis (right).For validation, the arginine finger which effects hydrolysis only, and Asp272 that coordinates GTP that effects nucleotide exchange were mutated (green box). Furthermore, residues at the Ras-like/all-alpha interface were investigated (blue box). They play a role in nucleotide exchange.
Figure 5:

Multiple turnover kinetics of GDP to GTP exchange and GTP hydrolysis under steady-state conditions in Gαi1 (left), and the underlying kinetics of single turnover nucleotide exchange (middle) and single turnover GTP hydrolysis (right).

For validation, the arginine finger which effects hydrolysis only, and Asp272 that coordinates GTP that effects nucleotide exchange were mutated (green box). Furthermore, residues at the Ras-like/all-alpha interface were investigated (blue box). They play a role in nucleotide exchange.

Single turnover measurements of the RGS4-catalyzed GTPase reaction in Gαi1 using time-resolved FTIR spectroscopy revealed an acceleration of about two orders of magnitude, as compared to the intrinsic reaction. Further integration of these data with biomolecular simulations decoded the IR-spectra and provided an RGS4 catalysis mechanism at subatomic resolution. In particular, the role of the catalytic intrinsic arginine finger was elucidated and found to be in a similar position as the arginine finger provided externally by Ras·GAP. The most striking feature is the different role of the intrinsic arginine finger with, and without, the RGS protein. In the X-ray structural models, the nucleotide is substituted by GTPγS or GDP·AlFx (Coleman et al., 1994; Kaya et al., 2016), and the position of the arginine is artificially altered due to the presence of the γS and AlF4 analogs. In contrast to this, QM/MM calculations of Gαi1 with the natural nucleotide GTP show a direct hydrogen bond to the γ-phosphate (Figure 6A) (Mann et al., 2016). This interaction was further confirmed experimentally by FTIR spectroscopy: A mutation deleting this hydrogen bond leads to a blue-shift of the absorption band of the γ-phosphate only, whereas the α- and β-phosphate bands are unchanged. Unlike Ras·GAP, RGS has no direct contact with the phosphate groups. Integration of QM/MM calculations and FTIR spectroscopy further revealed that RGS catalysis is achieved through the intrinsic arginine finger of Gαi1. RGS pushes the arginine finger into a bidentate coordination of the GTP with two additional hydrogen bonds to the α-phosphate (Figure 6B). A similar effect of the arginine finger on the α-phosphate is also seen in Ras. The bidentate coordination shifts the terminal PO3 of the γ-phosphate to a more planar conformation, closer to the transition state of the nucleophilic attack of water at the γ-phosphate. Notably, the planar structure of the Gαi1 transition state is also observed with the analog GDP·AlFx (Tesmer et al., 1997). Further, as in Ras, the negative charge is transferred towards the β-GTP, inducing a more product-like charge distribution. Whereas in GTP the formal charge of the β-GTP is −1, for β-GDP, it is −2. The more dissociative the mechanism, the closer the charge distribution of the transition state is to that of the product (Carvalho et al., 2015). The α-phosphate is turned into an eclipsed conformation relative to the β-GTP via close coordination between the arginine finger and Thr48. As similarly noted for the Ras·GAP system, this induces strain that can facilitate bond breakage (Rudack et al., 2012).

Figure 6: The role of the intrinsic arginine finger of Gαi1 with, and without, RGS4.
Figure 6:

The role of the intrinsic arginine finger of Gαi1 with, and without, RGS4.

An advantage of the simultaneous investigation of small and heterotrimeric GTPases is that the presence of different rate-limiting steps can help to elucidate common mechanisms, which are not resolved in the other system. For example, although it is not possible to measure coupling of the arginine finger to GTP in the Ras-GAP system, this was nicely resolved in the Gαi1 system (Mann et al., 2016). Similarly, mutation of Lys16, which is found in the conserved GTPase GxxxxGKS/T motif, was not possible in small GTPases due to the fact that mutated proteins were unstable (Du and Sprang, 2009). However, this mutation could be generated in Gαi1, because the binding pocket is more closed to the outside. Measurement of FTIR spectra for this mutant, which were subsequently decoded by QM/MM calculations, elucidated the role of the conserved lysine. Specifically, this residue shifts negative charges from γ-GTP to β-GTP and thereby facilitates GTP hydrolysis similarly to the arginine finger, with the mutant showing similarly slowed hydrolysis kinetics (Mann et al., 2017). In conclusion, these findings reveal novel intrinsic and GAP-catalyzed hydrolysis mechanisms for G-proteins at subatomic resolution (Figure 7).

Figure 7: Schematic mechanism for GTP hydrolysis in Ras and Gαi1.The most important residues for GTP hydrolysis (red or green) and nucleotide exchange (yellow) are given. The conserved lysine and the arginine finger transfer charges toward β-GTP and thereby facilitate GTP hydrolysis. The arginine finger induces a twist at α-GTP, in case of Gαi1 aided by Thr48. Further charge transfer is caused by Mg2+, which is precisely positioned via a conserved aspartic acid. Furthermore, the nucleophilic attacking water molecule is held in position by a conserved glutamine. Residues that affect the nucleotide exchange, either by direct binding or allosterically, are highlighted in yellow. Charge shifts and geometry twisting induce strain in GTP, facilitating GTP hydrolysis.
Figure 7:

Schematic mechanism for GTP hydrolysis in Ras and Gαi1.

The most important residues for GTP hydrolysis (red or green) and nucleotide exchange (yellow) are given. The conserved lysine and the arginine finger transfer charges toward β-GTP and thereby facilitate GTP hydrolysis. The arginine finger induces a twist at α-GTP, in case of Gαi1 aided by Thr48. Further charge transfer is caused by Mg2+, which is precisely positioned via a conserved aspartic acid. Furthermore, the nucleophilic attacking water molecule is held in position by a conserved glutamine. Residues that affect the nucleotide exchange, either by direct binding or allosterically, are highlighted in yellow. Charge shifts and geometry twisting induce strain in GTP, facilitating GTP hydrolysis.

Membrane binding of Ras

Ras is a peripheral membrane protein that is attached to the membrane by lipid anchors. Four isoforms (H-Ras, N-Ras, K-Ras-4A, and K-Ras-4B) each contain a distinct anchor, which regulates membrane targeting (Rocks et al., 2005; Brunsveld et al., 2006; Ahearn et al., 2011; Tsai et al., 2015), and this specific Ras localization is also an essential factor for signaling. Similar to the way in which GDP-dissociation inhibitors (GDIs) influence Rho and Rab function, Ras localization can be influenced by the prenyl-binding protein PDEδ (Chandra et al., 2012), which binds to Ras, preventing membrane binding. Localization can also be altered via the cleavage of one (N-Ras, K-Ras-4A) or two (H-Ras) lipid anchors or by phosphorylation (K-Ras-4B) (Ahearn et al., 2011; Schmick et al., 2015).

It was shown more than 10 years ago that Ras can form nanoclusters (Hancock and Parton, 2005), and recently, we demonstrated dimerization of N-Ras at a 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) model membrane, suggesting that Ras plays an active role in this nanoclustering (Güldenhaupt et al., 2012). Specifically, through the integration of attenuated total reflection (ATR)-FTIR spectroscopy, MD simulations, and fluorescence resonance energy transfer (FRET) experiments, we showed for the first time that N-Ras dimerizes at the membrane (Figure 8), and several investigations on Ras dimerization have been published since then. For example, the Groves group found that H-Ras forms dimers on membrane surfaces (Lin et al., 2014); however, these dimers may be induced by a photochemical reaction within the confocal fluorescence microscope (Chung et al., 2016). Evidence for dimer formation by K-Ras was also obtained using super-resolution fluorescence microscopy (Nan et al., 2013, 2015), even though the distance between the monomers remains unclear and the dimerization might be influenced by the label, i.e. mCherry alone tethered to the membrane with the K-Ras tail behaved in exactly the same way as did mCherry-KRas. However, these studies showed that Ras dimerization activates the MAPK pathway and therefore likely plays an important role in signal transduction. Similarly, a role for dimerization and nanocluster formation in signaling (Solman et al., 2015; Roob et al., 2016), possibly due to orientation changes of the G-domain (Mazhab-Jafari et al., 2015), was also reported by other groups. Specifically, Muratcioglu et al. (2015) reported that K-Ras dimer formation is nucleotide dependent using dynamic light scattering and nuclear magnetic resonance (NMR) spectroscopy. Of note, this dimer formation by the Ras proteins provides a novel drug target (Santos, 2014; Stephen et al., 2014; Chen et al., 2016; Lu et al., 2016), and elucidating the dimer interface at atomic detail will facilitate therapeutic approaches that modulate this interaction.

Figure 8: N-Ras dimers at a POPC model membrane.(A) The N-Ras dimer with an interface consisting of residues from the β2, β3 loop, helix 4, and helix 5 as obtained by MD simulations (Güldenhaupt et al., 2012). The distance between the fluorophores in the simulations agrees nicely with the experimental FRET-determined distance. (B) The proposed dimer interface with coordinating residues as obtained by MD simulations. A similar interface was proposed for K-Ras (Spencer-Smith et al., 2016; Prakash et al., 2017).
Figure 8:

N-Ras dimers at a POPC model membrane.

(A) The N-Ras dimer with an interface consisting of residues from the β2, β3 loop, helix 4, and helix 5 as obtained by MD simulations (Güldenhaupt et al., 2012). The distance between the fluorophores in the simulations agrees nicely with the experimental FRET-determined distance. (B) The proposed dimer interface with coordinating residues as obtained by MD simulations. A similar interface was proposed for K-Ras (Spencer-Smith et al., 2016; Prakash et al., 2017).

In 2012, we proposed a model for the N-Ras dimer interface based on our experimental findings (Güldenhaupt et al., 2012). Subsequently, several Ras dimer interfaces, both similar and different to ours, were proposed based on MD simulations, mainly for K-Ras (Jang et al., 2016; Prakash et al., 2016; Sayyed-Ahmad et al., 2016), and currently, there is no consensus on the detailed dimer interface. Some proposals are questionable from a physiological point of view; for example, dimerization including the effector domain (Muratcioglu et al., 2015) would not lead to the observed increased signaling (Nan et al., 2013, 2015). While measurements in cells have revealed that Ras forms nanoclusters and/or dimers, and FRET between Ras molecules is in agreement with this (Solman et al., 2015), a detailed investigation of the interaction remains challenging. In addition, proteins like galectin seem to interact with Ras and play a role in nanocluster formation in the cell (Prior et al., 2003). However, it remains unclear whether this is primarily due to the general effects of galectin on the membrane or to a specific galectin-Ras interaction. Notably, a recent report describes the direct interaction of galectin-1 dimers with Raf effectors, indicating an indirect interaction with Ras (Blaževitš et al., 2016). Further, the GEF SOS binds two Ras molecules (Margarit et al., 2003; Sondermann et al., 2004; Gureasko et al., 2008), and the effector BRAF dimerizes and also potentially binds to two Ras molecules (Poulikakos et al., 2010). In both cases, the two Ras molecules are in a distance enabling FRET without any direct Ras-Ras interaction. Consequently, it is difficult to prove direct Ras-Ras interactions in cells. Recently, an NMR study did not detect dimerization using the G-domain without the anchor region and without a bilayer (Kovrigina et al., 2015). This indicates that interactions within the hypervariable region (HVR) and/or membrane attachment anchor are required for dimerization. The latter can possibly promote a preorientation that leads to an entropic benefit for dimerization, similar to a crosslink (Zaman et al., 2002). On the other hand, dynamic light scattering experiments indicate dimerization for the Ras G-domain alone when bound to GTP-γ-S (Muratcioglu et al., 2015). As H-, N- and K-Ras have differences in their helix 4 and helix 5 regions, it is possible that these lead to the variable behaviors observed for the different isoforms (Parker and Mattos, 2015). Thus, the unambiguous characterization of the correct dimerization interface is still challenging.

In summary, integration of FTIR difference spectroscopy and biomolecular simulations provides a detailed insight into Ras and Gαi1 catalysis by their respective GAPs via protein-protein interactions and common features have been elucidated. Using the ATR technique further allows us to study the role of membrane binding of Ras. This opens the door for reconstitution experiments at the membrane in a near native but accurately defined system.

Acknowledgments

This work was supported by the Deutsche Forschungsgemeinschaft within the SFB 642. Excellent, inspiring and very fruitful collaboration within the SFB 642 especially with Roger S. Goody, Alfred Wittinghofer and Herbert Waldmann is very gratefully acknowledged.

References

Ahearn, I.M., Haigis, K., Bar-Sagi, D., and Philips, M.R. (2011). Regulating the regulator: post-translational modification of RAS. Nat. Rev. Mol. Cell. Biol. 13, 39–51.10.1038/nrm3255Search in Google Scholar PubMed PubMed Central

Blaževitš, O., Mideksa, Y.G., Šolman, M., Ligabue, A., Ariotti, N., Nakhaeizadeh, H., Fansa, E.K., Papageorgiou, A.C., Wittinghofer, A., Ahmadian, M.R., et al. (2016). Galectin-1 dimers can scaffold Raf-effectors to increase H-ras nanoclustering. Sci. Rep. 6, 24165.10.1038/srep24165Search in Google Scholar PubMed PubMed Central

Boriack-Sjodin, P.A., Margarit, S.M., Bar-Sagi, D., and Kuriyan, J. (1998). The structural basis of the activation of Ras by Sos. Nature 394, 337–343.10.1038/28548Search in Google Scholar PubMed

Brunsveld, L., Kuhlmann, J., Alexandrov, K., Wittinghofer, A., Goody, R.S., and Waldmann, H. (2006). Lipidated Ras and Rab peptides and proteins-synthesis, structure, and function. Angew. Chem. Int. Ed. 45, 6622–6646.10.1002/anie.200600855Search in Google Scholar PubMed

Carvalho, A.T.P., Szeler, K., Vavitsas, K., Åqvist, J., and Kamerlin, S.C.L. (2015). Modeling the mechanisms of biological GTP hydrolysis. Arch. Biochem. Biophys. 582, 80–90.10.1016/j.abb.2015.02.027Search in Google Scholar PubMed

Cepus, V., Scheidig, A.J., Goody, R.S., and Gerwert, K. (1998). Time-resolved FTIR studies of the GTPase reaction of H-ras p21 reveal a key role for the β-phosphate. Biochemistry 37, 10263–10271.10.1021/bi973183jSearch in Google Scholar PubMed

Chandra, A., Grecco, H.E., Pisupati, V., Perera, D., Cassidy, L., Skoulidis, F., Ismail, S.A., Hedberg, C., Hanzal-Bayer, M., Venkitaraman, A.R., et al. (2012). The GDI-like solubilizing factor PDEδ sustains the spatial organization and signalling of Ras family proteins. Nat. Cell Biol. 14, 329.10.1038/ncb2462Search in Google Scholar

Chen, M., Peters, A., Huang, T., Nan, X. (2016). Ras dimer formation as a new signaling mechanism and potential cancer therapeutic target. Mini Rev. Med. Chem. 16, 391–403.10.2174/1389557515666151001152212Search in Google Scholar PubMed PubMed Central

Cherfils, J., Zeghouf, M. (2013). Regulation of small GTPases by GEFs, GAPs, and GDIs. Physiol. Rev. 93, 269–309.10.1152/physrev.00003.2012Search in Google Scholar PubMed

Chung, J.K., Lee, Y.K., Lam, H.Y.M., and Groves, J.T. (2016). Covalent Ras Dimerization on Membrane Surfaces through Photosensitized Oxidation. J. Am. Chem. Soc. 138, 1800–1803.10.1021/jacs.5b12648Search in Google Scholar PubMed PubMed Central

Coleman, D.E., Berghuis, A.M., Lee, E., Linder, M.E., Gilman, A.G., and Sprang, S.R. (1994). Structures of active conformations of Gi alpha 1 and the mechanism of GTP hydrolysis. Science 265, 1405–1412.10.1126/science.8073283Search in Google Scholar PubMed

Cox, A.D. and Der, C.J. (2010). Ras history – the saga continues. Small GTPases 1, 2–27.10.4161/sgtp.1.1.12178Search in Google Scholar PubMed PubMed Central

Daumke, O. and Praefcke, G.J.K. (2016). Mechanisms of GTP hydrolysis and conformational transitions in the dynamin superfamily: mechanisms of GTP hydrolysis in the dynamin superfamily. Biopolymers 105, 580–593.10.1002/bip.22855Search in Google Scholar PubMed PubMed Central

Du, X. and Sprang, S.R. (2009). Transition state structures and the roles of catalytic residues in GAP-facilitated GTPase of Ras as elucidated by 18O kinetic isotope effects. Biochemistry 48, 4538–4547.10.1021/bi802359bSearch in Google Scholar PubMed PubMed Central

Dumitrescu, C.E. and Collins, M.T. (2008). McCune-Albright syndrome. Orphanet J. Rare. Dis. 3, 12.10.1186/1750-1172-3-12Search in Google Scholar PubMed PubMed Central

Gideon, P., John, J., Frech, M., Lautwein, A., Clark, R., Scheffler, J.E., and Wittinghofer, A. (1992). Mutational and kinetic analyses of the GTPase-activating protein (GAP)-p21 interaction: the C-terminal domain of GAP is not sufficient for full activity. Mol. Cell. Biol. 12, 2050–2056.10.1128/mcb.12.5.2050-2056.1992Search in Google Scholar

Güldenhaupt, J., Rudack, T., Bachler, P., Mann, D., Triola, G., Waldmann, H., Kötting, C., and Gerwert, K. (2012). N-Ras forms dimers at POPC membranes. Biophys. J. 103, 1585–1593.10.1016/j.bpj.2012.08.043Search in Google Scholar PubMed PubMed Central

Gureasko, J., Galush, W.J., Boykevisch, S., Sondermann, H., Bar-Sagi, D., Groves, J.T., and Kuriyan, J. (2008). Membrane-dependent signal integration by the Ras activator Son of sevenless. Nat. Struct. Mol. Biol. 15, 452–461.10.1038/nsmb.1418Search in Google Scholar PubMed PubMed Central

Gureasko, J., Kuchment, O., Makino, D.L., Sondermann, H., Bar-Sagi, D., and Kuriyan, J. (2010). Role of the histone domain in the autoinhibition and activation of the Ras activator Son of Sevenless. Proc. Natl. Acad. Sci. USA 107, 3430–3435.10.1073/pnas.0913915107Search in Google Scholar PubMed PubMed Central

Hamm, H.E. (1998). The many faces of G protein signaling. J. Biol. Chem. 273, 669–672.10.1074/jbc.273.2.669Search in Google Scholar PubMed

Hancock, J.F. and Parton, R.G. (2005). Ras plasma membrane signalling platforms. Biochem. J. 389, 1–11.10.1042/BJ20050231Search in Google Scholar PubMed PubMed Central

Jamali, T., Jamali, Y., Mehrbod, M., and Mofrad, M.R.K. (2011). Nuclear pore complex: biochemistry and biophysics of nucleocytoplasmic transport in health and disease. Int. Rev. Cell Mol. Biol. 287, 233–286.10.1016/B978-0-12-386043-9.00006-2Search in Google Scholar PubMed

Jang, H., Muratcioglu, S., Gursoy, A., Keskin, O., and Nussinov, R. (2016). Membrane-associated Ras dimers are isoform-specific: K-Ras dimers differ from H-Ras dimers. Biochem. J. 473, 1719–1732.10.1042/BCJ20160031Search in Google Scholar PubMed PubMed Central

Jencks, W.P. (2006) Binding energy, specificity, and enzymic catalysis: the circe effect. In: Advances in Enzymology – and Related Areas of Molecular Biology. A. Meister. ed. (Hoboken, NJ, USA: John Wiley & Sons, Inc.), pp. 219–410.10.1002/9780470122884.ch4Search in Google Scholar

Kaya, A.I., Lokits, A.D., Gilbert, J.A., Iverson, T.M., Meiler, J., and Hamm, H.E. (2016). A conserved hydrophobic core in Gαi1 regulates G protein activation and release from activated receptor. J. Biol. Chem. 291, 19674–19686.10.1074/jbc.M116.745513Search in Google Scholar PubMed PubMed Central

Khan, A.R. and Ménétrey, J. (2013). Structural biology of Arf and Rab GTPases’ effector recruitment and specificity. Structure 21, 1284–1297.10.1016/j.str.2013.06.016Search in Google Scholar PubMed

Kötting, C. and Gerwert, K. (2004). Time-resolved FTIR studies provide activation free energy, activation enthalpy and activation entropy for GTPase reactions. Chem. Phys. 307, 227–232.10.1016/j.chemphys.2004.06.051Search in Google Scholar

Kötting, C. and Gerwert, K. (2015). What vibrations tell us about GTPases. Biol. Chem. 396, 131–144.10.1515/hsz-2014-0219Search in Google Scholar PubMed

Kötting, C., Blessenohl, M., Suveyzdis, Y., Goody, R.S., Wittinghofer, A., and Gerwert, K. (2006). A phosphoryl transfer intermediate in the GTPase reaction of Ras in complex with its GTPase-activating protein. Proc. Nat. Acad. Sci. USA 103, 13911–13916.10.1073/pnas.0604128103Search in Google Scholar PubMed PubMed Central

Kötting, C., Kallenbach, A., Suveyzdis, Y., Wittinghofer, A., and Gerwert, K. (2008). The GAP arginine finger movement into the catalytic site of Ras increases the activation entropy. Proc. Nat. Acad. Sci. USA 105, 6260–6265.10.1073/pnas.0712095105Search in Google Scholar PubMed PubMed Central

Kovrigina, E.A., Galiakhmetov, A.R., and Kovrigin, E.L. (2015). The Ras G domain lacks the intrinsic propensity to form dimers. Biophys. J. 109, 1000–1008.10.1016/j.bpj.2015.07.020Search in Google Scholar PubMed PubMed Central

Li, G. and Zhang, X.C. (2004). GTP hydrolysis mechanism of Ras-like GTPases. J. Mol. Biol. 340, 921–932.10.1016/j.jmb.2004.06.007Search in Google Scholar PubMed

Ligeti, E., Welti, S., and Scheffzek, K. (2012). Inhibition and termination of physiological responses by GTPase activating proteins. Physiol. Rev. 92, 237–272.10.1152/physrev.00045.2010Search in Google Scholar

Lin, W-C., Iversen, L., Tu, H.-L., Rhodes, C., Christensen, S.M., Iwig, J.S., Hansen, S.D., Huang, W.Y., Groves, J.T. (2014). H-Ras forms dimers on membrane surfaces via a protein-protein interface. Proc. Natl. Acad. Sci. USA 111, 2996–3001.10.1073/pnas.1321155111Search in Google Scholar

Linder, M.E., Ewald, D.A., Miller, R.J., and Gilman, A.G. (1990). Purification and characterization of Go α and three types of Gi α after expression in Escherichia coli. J. Biol. Chem. 265, 8243–8251.10.1016/S0021-9258(19)39064-7Search in Google Scholar

Lu, S., Jang, H., Muratcioglu, S, Gursoy, A., Keskin, O., Nussinov, R., and Zhang, J. (2016). Ras conformational ensembles, allostery, and signaling. Chem. Rev. 116, 6607–6665.10.1021/acs.chemrev.5b00542Search in Google Scholar

Malbon, C.C. (2005). Developmental cell biology: G proteins in development. Nat. Rev. Mol. Cell. Biol. 6, 689–701.10.1038/nrm1716Search in Google Scholar

Mann, D., Teuber, C., Tennigkeit, S.A., Schröter, G., Gerwert, K., and Kötting, C. (2016). Mechanism of the intrinsic arginine finger in heterotrimeric G proteins. Proc. Natl. Acad. Sci. USA 113, E8041–E8050.10.1073/pnas.1612394113Search in Google Scholar

Mann, D., Höweler, U., Kötting, C., and Gerwert, K. (2017). Elucidation of single hydrogen bonds in GTPases via experimental and theoretical infrared spectroscopy. Biophys. J. 112, 66–77.10.1016/j.bpj.2016.11.3195Search in Google Scholar

Maracci, C. and Rodnina, M.V. (2016). Translational GTPases. Biopolymers 105, 463–475.10.1002/bip.22832Search in Google Scholar

Margarit, S.M., Sondermann, H., Hall, B.E., Nagar, B., Hoelz, A., Pirruccello, M., Bar-Sagi, D., and Kuriyan, J. (2003). Structural evidence for feedback activation by Ras.GTP of the Ras-specific nucleotide exchange factor SOS. Cell 112, 685–95.10.1016/S0092-8674(03)00149-1Search in Google Scholar

Mazhab-Jafari, M.T., Marshall, C.B., Smith, M.J., Gasmi-Seabrook, G.M., Stathopulos, P.B., Inagaki, F., Kay, L.E., Neel, B.G., and Ikura, M. (2015). Oncogenic and RASopathy-associated K-RAS mutations relieve membrane-dependent occlusion of the effector-binding site. Proc. Natl. Acad. Sci. USA 112, 6625–6630.10.1073/pnas.1419895112Search in Google Scholar PubMed PubMed Central

Mishra, A.K. and Lambright, D.G. (2016). Small GTPases and their GAPs. Biopolymers 105, 431–448.10.1002/bip.22833Search in Google Scholar PubMed PubMed Central

Muratcioglu, S., Chavan, T.S., Freed, B.C., Jang, H., Khavrutskii, L., Freed, R.N., Dyba, M.A., Stefanisko, K., Tarasov, S.G., Gursoy, A., et al. (2015). GTP-dependent K-Ras dimerization. Structure 23, 1325–1335.10.1016/j.str.2015.04.019Search in Google Scholar PubMed PubMed Central

Nan, X., Collisson, E.A., Lewis, S., Huang, J., Tamgüney, T.M., Liphardt, J.T., McCormick, F., Gray, J.W., and Chu, S. (2013). Single-molecule superresolution imaging allows quantitative analysis of RAF multimer formation and signaling. Proc. Natl. Acad. Sci. USA 110, 18519–18524.10.1073/pnas.1318188110Search in Google Scholar PubMed PubMed Central

Nan, X., Tamgüney, T.M., Collisson, E.A., Lin, L.J., Pitt, C., Galeas, J., Lewis, S., Gray, J.W., McCormick, F., and Chu, S. (2015). Ras-GTP dimers activate the mitogen-activated protein kinase (MAPK) pathway. Proc. Natl. Acad. Sci. USA 112, 7996–8001.10.1073/pnas.1509123112Search in Google Scholar PubMed PubMed Central

Neal, S.E., Eccleston, J.F., and Webb, M.R. (1990). Hydrolysis of GTP by p21NRAS, the NRAS protooncogene product, is accompanied by a conformational change in the wild-type protein: use of a single fluorescent probe at the catalytic site. Proc. Natl. Acad. Sci. USA 87, 3562–3565.10.1073/pnas.87.9.3562Search in Google Scholar PubMed PubMed Central

Northup, J.K., Sternweis, P.C., Smigel, M.D., Schleifer, L.S., Ross, E.M., and Gilman, A.G. (1980). Purification of the regulatory component of adenylate cyclase. Proc. Natl. Acad. Sci. USA 77, 6516–6520.10.1073/pnas.77.11.6516Search in Google Scholar PubMed PubMed Central

O’Hayre, M., Vázquez-Prado, J., Kufareva, I., Stawiski, E.W., Handel, T.M., Seshagiri, S., Gutkind, J.S. (2013). The emerging mutational landscape of G proteins and G-protein-coupled receptors in cancer. Nat. Rev. Cancer 13, 412–424.10.1038/nrc3521Search in Google Scholar PubMed PubMed Central

Pai, E.F. (1990). Refined crystal structure of the triphosphate conformation of H-ras p21 at 1.35 Å resolution: implications for the mechanism of GTP hydrolysis. EMBO J. 9, 2351–9.10.1002/j.1460-2075.1990.tb07409.xSearch in Google Scholar

Parker, J.A. and Mattos, C. (2015). The Ras-membrane interface: isoform-specific differences in the catalytic domain. Mol. Cancer Res. 13, 595–603.10.1158/1541-7786.MCR-14-0535Search in Google Scholar PubMed

Poulikakos, P.I., Zhang, C., Bollag, G., Shokat, K.M., Rosen, N. (2010). RAF inhibitors transactivate RAF dimers and ERK signalling in cells with wild-type BRAF. Nature 464, 427–430.10.1038/nature08902Search in Google Scholar PubMed PubMed Central

Prakash, P., Zhou, Y., Liang, H., Hancock, J.F., Gorfe, A.A. (2016). Oncogenic K-Ras binds to an anionic membrane in two distinct orientations: a molecular dynamics analysis. Biophys. J. 110, 1125–1138.10.1016/j.bpj.2016.01.019Search in Google Scholar PubMed PubMed Central

Prakash, P., Sayyed-Ahmad, A., Cho, K.-J., Dolino, D.M., Chen, W., Li, H., Grant, B.J., Hancock, J.F., and Gorfe, A.A. (2017). Computational and biochemical characterization of two partially overlapping interfaces and multiple weak-affinity K-Ras dimers. Sci. Rep. 7, 40109.10.1038/srep40109Search in Google Scholar PubMed PubMed Central

Prior, I.A., Muncke, C., Parton, R.G., and Hancock, J.F. (2003). Direct visualization of Ras proteins in spatially distinct cell surface microdomains. J. Cell Biol. 160, 165–170.10.1083/jcb.200209091Search in Google Scholar PubMed PubMed Central

Pylayeva-Gupta, Y., Grabocka, E., and Bar-Sagi, D. (2011). RAS oncogenes: weaving a tumorigenic web. Nat. Rev. Cancer 11, 761–774.10.1038/nrc3106Search in Google Scholar PubMed PubMed Central

Rasmussen, S.G., DeVree, B.T., Zou, Y., Kruse, A.C., Chung, K.Y., Kobilka, T.S., Thian, F.S., Chae, P.S., Pardon, E., Calinski, D., et al. (2011). Crystal structure of the β2 adrenergic receptor-Gs protein complex. Nature 477, 549–55.10.1038/nature10361Search in Google Scholar PubMed PubMed Central

Rittinger, K., Walker, P.A., Eccleston, J.F., Smerdon, S.J., and Gamblin, S.J. (1997a). Structure at 1.65 Å of RhoA and its GTPase-activating protein in complex with a transition-state analogue. Nature 389, 758–762.10.1038/39651Search in Google Scholar PubMed

Rittinger, K., Walker, P.A., Eccleston, J.F., Nurmahomed, K., Owen, D., Laue, E., Gamblin, S.J., and Smerdon, S.J. (1997b). Crystal structure of a small G protein in complex with the GTPase-activating protein rhoGAP. Nature 388, 693–697.10.1038/41805Search in Google Scholar PubMed

Rocks, O., Peyker, A., Kahms, M., Verveer, P.J., Koerner, C., Lumbierres, M., Kuhlmann, J., Waldmann, H., Wittinghofer, A., and Bastiaens, P.I. (2005). An acylation cycle regulates localization and activity of palmitoylated Ras isoforms. Science 307, 1746–1752.10.1126/science.1105654Search in Google Scholar PubMed

Roob, E., Trendel, N., Rein ten Wolde, P., and Mugler, A. (2016). Cooperative clustering digitizes biochemical signaling and enhances its fidelity. Biophys. J. 110, 1661–1669.10.1016/j.bpj.2016.02.031Search in Google Scholar PubMed PubMed Central

Roston, D. and Cui, Q. (2016). Substrate and transition state binding in alkaline phosphatase analyzed by computation of oxygen isotope effects. J. Am. Chem. Soc. 138, 11946–11957.10.1021/jacs.6b07347Search in Google Scholar PubMed PubMed Central

Rudack, T., Xia, F., Schlitter, J., Kötting, C., and Gerwert, K. (2012). Ras and GTPase-activating protein (GAP) drive GTP into a precatalytic state as revealed by combining FTIR and biomolecular simulations. Proc. Natl. Acad. Sci. USA 109, 15295–15300.10.1073/pnas.1204333109Search in Google Scholar PubMed PubMed Central

Santos, E. (2014). Dimerization opens new avenues into Ras signaling research. Sci. Signal. 7, pe12–pe12.10.1126/scisignal.2005318Search in Google Scholar PubMed

Sayyed-Ahmad, A., Cho, K.-J., Hancock, J.F., and Gorfe, A.A. (2016). Computational equilibrium thermodynamic and kinetic analysis of K-ras dimerization through an effector binding surface suggests limited functional role. J. Phys. Chem. B 120, 8547–8556.10.1021/acs.jpcb.6b02403Search in Google Scholar PubMed PubMed Central

Scheffzek, K., Ahmadian, M.R., Kabsch, W., Wiesmüller, L., Lautwein, A., Schmitz, F., and Wittinghofer, A. (1997). The Ras-RasGAP complex: structural basis for GTPase activation and its loss in oncogenic Ras mutants. Science 277, 333–8.10.1126/science.277.5324.333Search in Google Scholar PubMed

Schmick, M., Kraemer, A., and Bastiaens, P.I.H. (2015). Ras moves to stay in place. Trends Cell Biol. 25, 190–197.10.1016/j.tcb.2015.02.004Search in Google Scholar PubMed

Schröter, G., Mann, D., Kötting, C., and Gerwert, K. (2015). Integration of fourier transform infrared spectroscopy, fluorescence spectroscopy, steady-state kinetics and molecular dynamics simulations of Gαi1 distinguishes between the GTP hydrolysis and GDP Release Mechanism. J. Biol. Chem. 290, 17085–17095.10.1074/jbc.M115.651190Search in Google Scholar PubMed PubMed Central

Sit, S.-T. and Manser, E. (2011). Rho GTPases and their role in organizing the actin cytoskeleton. J. Cell. Sci. 124, 679–683.10.1242/jcs.064964Search in Google Scholar PubMed

Solman, M., Ligabue, A., Blazevits, O., Jaiswal, A., Zhou, Y., Liang, H., Lectez, B., Kopra, K., Guzmán, C., Härmä, H., et al. (2015). Specific cancer associated mutations in the switch III-region of Ras increase tumorigenicity by nanocluster augmentation. eLife 4, e08905.10.7554/eLife.08905Search in Google Scholar PubMed PubMed Central

Sondermann, H., Soisson, S.M., Boykevisch, S., Yang, S.S., Bar-Sagi, D., Kuriyan, J. (2004). Structural analysis of autoinhibition in the Ras activator son of sevenless. Cell 119, 393–405.10.1016/j.cell.2004.10.005Search in Google Scholar PubMed

Spencer-Smith, R., Koide, A., Zhou, Y., Eguchi, R.R., Sha, F., Gajwani, P., Santana, D., Gupta, A., Jacobs, M., Herrero-Garcia, E., et al. (2016). Inhibition of RAS function through targeting an allosteric regulatory site. Nat. Chem. Biol. 13, 62–68.10.1038/nchembio.2231Search in Google Scholar PubMed PubMed Central

Sprang, S.R. (2016). Activation of G proteins by GTP and the mechanism of Gα-catalyzed GTP hydrolysis: activation of G Proteins. Biopolymers 105, 449–462.10.1002/bip.22836Search in Google Scholar

Stenmark, H. (2009). Rab GTPases as coordinators of vesicle traffic. Nat. Rev. Mol. Cell. Biol. 10, 513–525.10.1038/nrm2728Search in Google Scholar

Stephen, A.G., Esposito, D., Bagni, R.K., McCormick, F. (2014). Dragging Ras back in the ring. Cancer Cell 25, 272–281.10.1016/j.ccr.2014.02.017Search in Google Scholar

Syrovatkina, V., Alegre, K.O., Dey, R., and Huang, X.-Y. (2016). Regulation, signaling, and physiological functions of G-Proteins. J. Mol. Biol. 428, 3850–3868.10.1016/j.jmb.2016.08.002Search in Google Scholar

Tesmer, J.J.G., Berman, D.M., Gilman, A.G., Sprang, S.R. (1997). Structure of RGS4 bound to AlF4 – activated G(i a1): stabilization of the transition state for GTP hydrolysis. Cell 89, 251–261.10.1016/S0092-8674(00)80204-4Search in Google Scholar

Tsai, F.D., Lopes, M.S., Zhou, M., Court, H., Ponce, O., Fiordalisi, J.J., Gierut, J.J., Cox, A.D., Haigis, K.M., and Philips, M.R. (2015). K-Ras4A splice variant is widely expressed in cancer and uses a hybrid membrane-targeting motif. Proc. Natl. Acad. Sci. USA 112, 779–784.10.1073/pnas.1412811112Search in Google Scholar PubMed PubMed Central

Tucker, J., Sczakiel, G., Feuerstein, J., John, J., Goody, R.S., and Wittinghofer, A. (1986). Expression of p21 proteins in Escherichia coli and stereochemistry of the nucleotide-binding site. EMBO J. 5, 1351–8.10.1002/j.1460-2075.1986.tb04366.xSearch in Google Scholar

Vetter, I.R. and Wittinghofer, A. (2001). Signal transduction – the guanine nucleotide-binding switch in three dimensions. Science 294, 1299–1304.10.1126/science.1062023Search in Google Scholar PubMed

Walland, A. (1975). cAMP as a second messenger in central blood pressure control. Naunyn-Schmiedeberg’s Arch. Pharmacol. 290, 419–423.10.1007/BF00499953Search in Google Scholar PubMed

Wennerberg, K., Rossman, K.L., and Der, C.J. (2005). The Ras superfamily at a glance. J. Cell Sci. 118, 843–846.10.1242/jcs.01660Search in Google Scholar PubMed

Wittinghofer, A. (2014a). Ras Superfamily Small G Proteins: Biology and Mechanisms 1 (Vienna: Springer Vienna).10.1007/978-3-7091-1806-1Search in Google Scholar

Wittinghofer, A. (2014b). Ras Superfamily Small G Proteins: Biology and Mechanisms 2 (Vienna: Springer Vienna).10.1007/978-3-319-07761-1Search in Google Scholar

Wittinghofer, A. and Vetter, I.R. (2011). Structure-function relationships of the G domain, a canonical switch motif. Annu. Rev. Biochem. 80, 943–971.10.1146/annurev-biochem-062708-134043Search in Google Scholar PubMed

Xia, F., Rudack, T., Cui, Q., Kotting, C., and Gerwert, K. (2012). Detailed structure of the H2PO4-guanosine diphosphate intermediate in Ras-GAP decoded from FTIR experiments by biomolecular simulations. J. Am. Chem. Soc. 134, 20041–20044.10.1021/ja310496eSearch in Google Scholar PubMed

Zaman, M.H., Berry, R.S., and Sosnick, T.R. (2002). Entropic benefit of a cross-link in protein association. Proteins 48, 341–351.10.1002/prot.10165Search in Google Scholar PubMed

Received: 2016-10-19
Accepted: 2017-2-15
Published Online: 2017-2-23
Published in Print: 2017-5-1

©2017, Klaus Gerwert et al., published by De Gruyter.

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/hsz-2016-0314/html
Scroll to top button