Next Article in Journal
Development and Characterization of an Oncolytic Human Adenovirus-Based Vector Co-Expressing the Adenovirus Death Protein and p14 Fusion-Associated Small Transmembrane Fusogenic Protein
Previous Article in Journal
Targeting Protein Misfolding and Aggregation as a Therapeutic Perspective in Neurodegenerative Disorders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Determinants for Guanine Binding in GTP-Binding Proteins: A Data Mining and Quantum Chemical Study

by
Pawan Bhatta
and
Xiche Hu
*
Department of Chemistry and Biochemistry, University of Toledo, Toledo, OH 43606, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(22), 12449; https://doi.org/10.3390/ijms252212449
Submission received: 1 November 2024 / Revised: 15 November 2024 / Accepted: 18 November 2024 / Published: 20 November 2024
(This article belongs to the Special Issue Latest Advances in Protein-Ligand Interactions)

Abstract

:
GTP-binding proteins are essential molecular switches that regulate a wide range of cellular processes. Their function relies on the specific recognition and binding of guanine within their binding pockets. This study aims to elucidate the molecular determinants underlying this recognition. A large-scale data mining of the Protein Data Bank yielded 298 GTP-binding protein complexes, which provided a structural foundation for a systematic analysis of the intermolecular interactions that are responsible for the molecular recognition of guanine in proteins. It was found that multiple modes of non-bonded interactions including hydrogen bonding, cation–π interactions, and π–π stacking interactions are employed by GTP-binding proteins for binding. Subsequently, the strengths of non-bonded interaction energies between guanine and its surrounding protein residues were quantified by means of the double-hybrid DFT method B2PLYP-D3/cc-pVDZ. Hydrogen bonds, particularly those involving the N2 and O6 atoms of guanine, confer specificity to guanine recognition. Cation–π interactions between the guanine ring and basic residues (Lys and Arg) provide significant electrostatic stabilization. π–π stacking interactions with aromatic residues (Phe, Tyr, and Trp) further contribute to the overall binding affinity. This synergistic interplay of multiple interaction modes enables GTP-binding proteins to achieve high specificity and stability in guanine recognition, ultimately underpinning their crucial roles in cellular signaling and regulation. Notably, the NKXD motif, while historically considered crucial for guanine binding in GTP-binding proteins, is not universally required. Our study revealed significant variability in hydrogen bonding patterns, with many proteins lacking the NKXD motif but still effectively binding guanine through alternative arrangements of interacting residues.

Graphical Abstract

1. Introduction

GTP-binding proteins, commonly known as G proteins, are critical regulators of a myriad of cellular processes, acting as molecular switches that facilitate signal transduction in response to extracellular stimuli [1,2,3,4]. Two of the most common GTP-binding proteins are heterotrimeric G proteins and small monomeric G proteins [5,6]. Heterotrimeric G proteins consist of three subunits—α, β, and γ—and are integral to signaling pathways mediated by G protein-coupled receptors (GPCRs). They can be further divided into families based on the α subunit, including the Gs family (which activates adenylyl cyclase), the Gi family (which inhibits adenylyl cyclase), the Gq family (which activates phospholipase C), and the G12/13 family (which regulates Rho family GTPases). Small monomeric G proteins, on the other hand, are single polypeptide chains that function independently in various signaling pathways. This group includes the Ras family, known for its role in cell proliferation and survival; the Rho family, which regulates cytoskeletal dynamics; the Rab family, involved in vesicular transport; the Arf family, associated with membrane trafficking; and the Ran family, crucial for nucleocytoplasmic transport. Other GTP-binding proteins include septins, tubulins, dynamins, eukaryotic translation initiation/elongation factors, etc.
Due to the important role of GTP-binding proteins in various cellular processes and diverse signaling transduction networks, the molecular recognition of guanine nucleotides (GTP, GDP, and GMP) in GTP-binding proteins has long been a subject of great interest. The concept of the molecular recognition of ligands in proteins has a historical origin based on Emil Fischer’s “lock-and-key” model and Daniel Koshland’s “induced fit” hypothesis. It is the advent of X-ray crystallography that enabled the ability to visualize the complex three-dimensional structures of GTP-binding proteins and their complexes, thereby elucidating the critical sequence motifs and binding sites integral to molecular recognition [7]. The sequence motif responsible for the binding of guanine to G proteins is G4, i.e., the N/TKXD sequence motif [8,9]. Hereafter, we will adopt the motif nomenclature of [8]. GTP-binding proteins also contain G1, i.e., the GXXXXGK(S/T) sequence motif (also known as the Walker A motif), which is responsible for binding the GTP’s phosphate group. This phosphate group is also present in ATP, and the same G1 sequence motif is responsible for the phosphate group recognition in the ATP-binding protein. The G2, i.e., X(T/S)X, and G3, i.e., DXXG (Walker B motif), sequence motifs are involved in the coordination of Mg2+ ions. G5 is the (T/G)(C/S)A sequence motif required to strengthen the guanine base recognition [8]. Multiple studies have been carried out to understand the molecular recognition of ribose sugar [10], a phosphate group, and its associated magnesium ion [11,12,13,14] of GTP and ATP in proteins. In this work, we aim to decipher the molecular determinants for molecular recognition of the guanine base in GTP-binding proteins.
It is now widely accepted that molecular recognition is mediated through non-covalent interactions (also known as non-bonded interactions) such as hydrogen bonding, metal coordination, van der Waals forces (VDW), cation–π interaction, π–π stacking interaction, CH–π interaction, XH–π interaction (X = N, O, S), salt bridge, etc. [15,16,17,18,19,20]. To understand the molecular recognition of guanine, one needs to know the specific non-bonded interactions between guanine and its surrounding residues in proteins.
Figure 1 shows the molecular structure of the guanine base, an aromatic motif that features multiple hydrogen bond acceptors and donors, which can form specific hydrogen bonds with the surrounding residues inside the GTP-binding pocket. It has the capacity to form as many as six hydrogen bonds, acting as a donor for three hydrogen bonds at the N1 and N2 positions, and hydrogen bond acceptors at the N3, O6, and N7 positions. This hydrogen bonding capacity of the guanine base is widely accepted as an important non-bonded interaction mode for DNA base-pairing and protein–ligand interactions. There are two more equally important non-bonded interaction modes for guanine–protein binding, i.e., π–π stacking interactions and cation–π interactions. Just as in the case of DNA base stacking, the conjugated π rings of the guanine base can interact with surrounding aromatic residues (Phe, Tyr, and Trp) via π–π stacking interactions. The conjugated π rings of guanine base can also interact with positively charged residues (Lys and Arg) through cation–π interactions. A wealth of information has been accumulated displaying the importance of π–π stacking interactions and cation–π interactions in the formation of biomolecular systems [15,16,20,21,22,23]. Typically, π–π stacking interactions and cation–π interactions are of similar or even greater magnitude than the hydrogen bonding energy, as shown by our ATP binding study [24] and other investigations of cation–π [25,26,27] and π–π stacking interactions [28,29,30] in biological systems. In our analysis of the molecular recognition of the guanine base of GTP reported here, π–π stacking interactions and cation–π interactions are systematically analyzed in addition to hydrogen bonding. Furthermore, contributions of each one of the non-bonded interaction modes (π–π stacking, cation–π interaction, and H-bonding) to binding between the guanine base and protein are quantified by means of high-level quantum chemical calculations. We are interested in determining which types of non-bonded interactions are used by GTP-binding proteins in the recognition of the guanine base and what their relative importance is.
In this study, the molecular determinants responsible for the molecular recognition of the guanine moiety in GTP-binding proteins were deciphered by means of data mining and high-level quantum chemical analysis. A large-scale data mining of the Protein Data Bank was carried out, which resulted in the establishment of a database of 298 nonredundant high-resolution GTP-binding proteins complexed with bound guanine nucleotides. For all these complexes, the modes of the non-bonded interactions between guanine and its surrounding residues were systematically analyzed to decipher the specific interactions responsible for molecular recognition. Furthermore, the contributions of each one of the non-bonded interaction modes (π–π stacking, cation–π interaction, and H-bonding) to binding between the guanine base and protein were quantified by means of high-level quantum chemical calculations.
The remainder of this article is structured as follows. The Results and Discussion section presents the binding environments of the guanine base in 298 guanylate-binding protein complexes, particularly focusing on the dominant hydrogen bonding patterns, the cation–π interactions between guanine bases, and the side chains of positively charged residues lysine and arginine, as well as the π–π stacking interactions between guanine bases and the side chains of aromatic residues such as phenylalanine, tyrosine, and tryptophan. The strengths of non-bonded interactions for the representatives of cation–π and π–π stacking interactions, as calculated using the B2PLYP-D3/cc-pVDZ method [31,32,33], are described. Then, a case study is detailed, illustrating the distribution of the energetic contributions from various modes of non-bonded interactions to the binding of guanine within the context of an entire protein. The biological significance of our findings is discussed at the end of this section. In the Theory and Methods section, we detail the procedures for data mining GTP-binding proteins from the Protein Data Bank (PDB) along with specific details regarding the B2PLYP-D3/cc-pVDZ electronic structure calculations of non-bonded interactions in the guanylate–protein complexes. A brief summary is provided in the Conclusion section.

2. Results and Discussion

The data mining analysis resulted in a total of 298 nonredundant high-resolution (2.5 Å or better) GTP-binding proteins complexed with bound guanine nucleotides. Proteins with over 90% sequence identity were excluded to minimize redundancy. Table 1 provides an extensive list of the protein complexes containing bound guanine nucleotides (GMP, GDP, and GTP), along with essential details such as the protein family to which each complex belongs, the nucleotide type, PDB IDs, and the structural resolution. These GTP-binding proteins belong to 57 protein families, which underscores the essential, multifaceted role of GTP in cellular biology. These diverse families cover a broad spectrum of biological functions: from basic cellular processes like protein synthesis (50S ribosome-binding GTPases and EF-Tu) and cell division (FtsZ and septins), to advanced regulatory systems in signaling (the Ras family and ARF family) and immune responses (Interferon-inducible GTPase and AIG1). Many families are directly involved in metabolic pathways (PEPCK and GTP cyclohydrolase I) or are structural and enzymatic scaffolds (MobA-like NTP Transferase and Ferrous Iron Transport), ensuring cellular health and adaptability. Moreover, the diverse spectrum of protein families provides a wide variety of guanine-binding pockets for our molecular recognition study.
Based on the three-dimensional structures, the binding pockets of the guanine bases within their respective target GTP-binding proteins were meticulously analyzed using the Visual Molecular Dynamics (VMD) program to identify residues that engage in non-bonded interactions with each guanine base. Consistent with the physical nature of each type of non-bonded interaction, a cut-off distance of 3.5 Å between the donor and the acceptor was used for hydrogen bonding, and a cut-off distance of 5.6 Å was used for π–π stacking and cation–π interactions. For the former interaction, a slightly longer distance of 3.5 Å, rather than the optimal hydrogen bonding range of 2.8 to 3.2 Å, was adopted to account for the dynamic fluctuations in atomic positions. For the latter interactions, the computed strengths of the solution-phase interaction energies typically diminished beyond 5.6 Å, as indicated by our prior quantum chemical calculations [22]. The non-bonded interactions (hydrogen bonding, π–π stacking, and cation–π interactions) so identified were carefully studied, with details tabulated in Table S1 of the Supplementary Materials. Those non-bonded interactions are described and analyzed below.

2.1. Hydrogen Bonding

Table 2 presents a summary of the hydrogen bond patterns between the guanine base and surrounding residues in GTP-binding proteins. The table categorizes the hydrogen bonding interactions based on sequence motifs and associated hydrogen-bonding patterns. An extensively detailed list of interaction mode of hydrogen bonds is given in Table S1.
We identified six distinct hydrogen bonding patterns that are employed by the surrounding residues of the GTP-binding proteins for the molecular recognition of the guanine base. Of these, four are associated with the NKXD motif, while the remaining two patterns lack this motif. For the former, we adopted a hydrogen bonding pattern notation based on the participation of the residues from the NKXD sequence motif in the hydrogen bond interactions. The one-letter residue code is colored red if the residue participates in hydrogen bonding with the guanine base. For clarity, each of the four hydrogen bonding patterns are illustrated with a representative example in Figure 2. Figure 2a depicts the Ni-Ki+1-Xi+2-Di+3 pattern, where the side chain of the asparagine (N), the main chain of lysine (K), and the side chain of aspartate (D) form multiple hydrogen bonds with guanine. Figure 2b illustrates the Ni-Ki+1-Xi+2-Di+3 pattern in which the side chain of the asparagine (N) and the side chain of aspartate (D) are involved in hydrogen bonding. Figure 2c shows the Ni-Ki+1-Xi+2-Di+3 pattern in which the main-chain amino group from lysine (K) and the side chain of aspartate (D) participate in hydrogen bonds. Figure 2d demonstrates the Ni-Ki+1-Xi+2-Di+3 pattern in which only the side chain of aspartate (D) participates in hydrogen bonding. The first Ni-Ki+1-Xi+2-Di+3 and the second Ni-Ki+1-Xi+2-Di+3 pattern both occur most frequently with a probability of 19.1%. The Ni-Ki+1-Xi+2-Di+3 pattern appears least frequently with a probability of 6.7%. The Ni-Ki+1-Xi+2-Di+3 pattern was observed in 11.7% of GTP-binding proteins.
Notably, the NKXD motif [9,35]—a fingerprint sequence for guanine-binding sites—emerges as a key player in forming hydrogen bonds with the guanine in these proteins. However, only 56.7% (169 out of 298) of the complexes that bind guanine have the NKXD sequence motif (see Table S1). Additionally, as described above, not all residues within the NKXD motif participate in hydrogen bond interactions.
Remarkably, in the remaining 43.3% of the GTP-binding proteins, the NKXD sequence motif is absent. Within this subset, 20.5% of GTP-binding proteins feature either an aspartate (D) or a glutamate (E) residue that forms hydrogen bonds with the hydrogen from the N1 or N2H1 atoms of the guanine base. We designate this interaction mode as the D/E plus motif, which is illustrated in Figure 2e. The remaining 22.8% of GTP-binding proteins lack a specific conserved sequence motif for guanine recognition, wherein any amino acid residue may form hydrogen bonds with at least one of the N1, N2, or O6 atoms of guanine, either directly or through structured water molecules. An example of the last pattern is given in Figure 2f.
From the perspective of the guanine base, it was found that hydrogen bond donors and acceptors at various positions of the guanine ring have different preferences for hydrogen bond formation. The N1 atom directly donates hydrogen bonds to the surrounding residues, which occurs in 87.3% (260 out of 298) GTP-binding proteins. The N2 atom, on the other hand, has the capacity to directly donate two hydrogens, N2H1 and N2H2, to the surrounding residues. In 89.6% (267 out of 298) of GTP-binding proteins, the N2H1 hydrogen is donated to surrounding residues. In 21.8% (65 out of 298) of GTP-binding proteins, the N2H2 hydrogen is donated to the surrounding residues. In many cases, the N2H2 is also donated to a nearby structured water molecule, which acts as a bridge for hydrogen bonding with amino acid residues in the GTP-binding pocket. It is worth noting that, in many cases, the hydrogen from N1 and N2H1 are donated to aspartate or glutamate to form a double-hydrogen bond (dual-hydrogen bond). Since many GTP-binding proteins have aspartate or glutamate in the GTP-binding pocket, this dual-hydrogen bond mode represents the dominant mode of hydrogen bond interaction. In 27.9% (83 out of 298) of GTP-binding proteins, the O6 atom accepts hydrogen bonds from the main chain amino group of at least one of the residues from the NKXD sequence motif. In 85.6% (255 out of 298) of GTP-binding proteins, O6 accepts an additional hydrogen bond from the main-chain amino group of at least alanine and/or its succeeding residue from the (T/G)(C/S)A sequence motif or any non-conserved residues. The N7 atom accepts a hydrogen bond from the surrounding residue in 55.4% (165 out of 298) of GTP-binding proteins, and the N3 atom accepts a hydrogen bond from the surrounding residues in 6.7% (20 out of 298) of GTP-binding proteins. Interestingly, it was observed that a conserved structured water molecule near the N3 atom donates the hydrogen bond in 41.3% (123 out of 298) of GTP-binding proteins.
Based on the above analysis, it is evident that the most frequent hydrogen bond participating atoms/groups in the guanine base are the N1, N2, and O6 atoms. The N3 and N7 atoms, also present in ATP, are less preferred for hydrogen bonding by proteins that bind guanine.

2.2. Cation–π Interaction

Cation–π interactions were systematically examined across all 298 GTP-binding proteins, revealing this interaction as a prevalent non-bonded interaction mode for GTP binding in proteins. In 86.6% of GTP-binding proteins (258 out of 298), at least one cation–π interaction does exist between the guanine base and positively charged side chains of the interacting residues. Furthermore, about 48% of the complexes have more than one positively charged residue interacting with guanine. These complexes are aligned by the superimposition of the guanine base, and Figure 3 shows a 3D stereo drawing of the aligned GTP-protein complexes featuring one or more positively charged residues (lysine and arginine) within 5.6 Å of the guanine base.
To quantitatively establish the contribution of cation–π interactions to the binding of guanine with its targeted proteins, the strengths of cation–π interactions between the guanine base and its interacting residues were quantified by means of quantum chemical calculations. For this purpose, 12 distinctive interacting intermolecular pairs between guanine and the positively charged residue were chosen based on Figure 3. They were selected according to two criteria: representation and uniqueness. The representative intermolecular pair is the pair that samples the most abundant regions of Figure 3. The unique intermolecular pair is the one that is uniquely situated in Figure 3 in terms of position and orientation. The three-dimensional structures for 9 of these 12 cation–π interacting pairs are presented in Figure 4. The strengths of the non-bonded interaction energies between guanine and its surrounding aromatic residues were quantified in a pairwise manner using the double-hybrid DFT method at the B2PLYP-D3/cc-pVDZ level of theory (see the Theory and Methods section for details). The resulting pairwise non-bonded interaction energies for the selected cation–π interactions are detailed in Table 3. The magnitudes of cation–π interactions are moderate to strong, ranging from −1.51 to −10.61 kcal/mol. These results indicate the vital role of cation–π interactions in stabilizing guanine binding within GTP-binding proteins.
What are the factors that control the strength of the intermolecular cation–π interaction? This is an important question we want to address below. The intermolecular distance was found to be the predominant factor determining the strength of the non-bonded interaction energy ( Δ E I n t a q ). As shown in Table 3, as the intermolecular distance increased, the interaction energy decreased. However, at the same time, the extent of overlap between the side chain of positively charged residue and the guanine ring also influenced the intermolecular interaction energy. The representative cation–π intermolecular pairs associated with PDB 1G7S, 1S4O, and 5A07 have the largest interaction energy (see Table 3 and Figure 4). As shown in Figure 4, these cation–π intermolecular pairs have the greatest extent of side chain overlap with the guanine ring. In some unique cation–π intermolecular pairs, a dual mode of non-bonded interactions is seen where cation–π and hydrogen bond interactions both simultaneously exist, e.g., the intermolecular pairs in 1RYA, 2IRX, and 6B9F. The interaction energies are much larger in these cases. This analysis leads us to conclude that the strength of the intermolecular cation–π interaction is dependent upon the combination of three factors, i.e., intermolecular distance, the extent of the side chain overlap of positively charged residues with guanine ring, and the existence of multiple modes of interaction (cation–π and hydrogen bond).

2.3. π–π Stacking Interaction

The binding pockets of the guanine bases in all 298 GTP-binding proteins were examined to identify the aromatic residues capable of π–π stacking interactions. In 54.4% of the complexes, or 162 out of 298, π–π stacking interaction does exist between the guanine base and aromatic side chains. Figure 5 displays all 162 of the GTP-binding proteins with aromatic residues within 5.6 Å of the guanine base. The aromatic residues Phe, Tyr, and Trp form four primary clusters surrounding the guanine bases: to the left and right, at the top, and at the bottom. The arrangements at the top and bottom are typically categorized as either parallel face-to-face stacking or parallel-displaced stacking, depending on the degree of displacement of the aromatic centers. When aromatic residues perpendicularly approach the π-plane of the guanine base, this configuration is referred to as a “T-shaped” edge-to-face arrangement. The observed distribution pattern of the aromatic residues around the π-plane of the guanine base appears to be optimal, as suggested by modeling studies of the benzene dimer, which is commonly regarded as the standard model for aromatic π–π stacking interactions [36].
To quantitatively assess the contribution of π–π stacking interactions to the binding of guanine with their target proteins, we employed quantum chemical calculations to evaluate the strength of these interactions between the guanine base and its interacting residues. We selected 14 distinct intermolecular pairs (comprising guanine and aromatic residues) based on the patterns shown in Figure 5. These pairs were chosen according to two criteria: representativeness and uniqueness. A representative pair samples the most prevalent regions in Figure 5, while a unique pair is specifically positioned and oriented within that figure. The three-dimensional structures of 6 of these 14 π–π stacking interacting pairs are illustrated in Figure 6.
The strengths of the non-bonded interaction energies between guanine and its surrounding aromatic residues were quantified in a pairwise manner using the double-hybrid DFT method at the B2PLYP-D3/cc-pVDZ level of theory (see the Theory and Methods section for further details). The resulting pairwise interaction energies for the selected π–π stacking interactions are presented in Table 4.
The π–π stacking interaction energies were found to be low to moderate, ranging from −0.34 to −6.57 kcal/mol. The intermolecular distance is the predominant factor determining the strength of the non-bonded interaction energy ( Δ E I n t a q ), as can be seen in Table 4 as the intermolecular distance increases the interaction energy decreases. However, at the same time, the angles between two interacting ring planes and the π–π stacking conformations also influence the non-bonded interaction energy. The interaction energy is larger in those cases where the angle between the ring’s planes is nearly zero and the rings are in almost parallel displaced configurations, e.g., the unique π–π stacking intermolecular pair in 1RYA, 3R4V, and 3DZH (see Table 4 and Figure 6). These findings illustrate that π–π stacking interactions, though generally weaker than cation–π interactions, contribute to the stability of the guanine binding in GTP-binding proteins.

2.4. Energetic Contribution by Various Modes of Non-Bonded Interactions to the Binding of Guanine in a Representative Complex

The distribution of modes of the non-bonded interactions in the GTP-binding proteins was systematically examined based on their X-ray crystal structures (see Table S1). The objective was to decipher the relative importance of the different modes of non-bonded interactions for the molecular recognition of the guanine base in proteins. Due to space limitation, we chose one GTP-binding protein, i.e., the p21-ras protein (PDB ID: 1QRA), as an illustration. One of the main reasons for the choice of 1QRA is its representativeness; it features the Ni-Ki+1-Xi+2-Di+3 hydrogen bond pattern (see Section 2.1 and Table S1). The latter, along with the other three patterns (see above) are associated with the major sequence motif NKXD. On the basis of the 1.6 Å resolution X-ray crystal structure [37] (PDB ID: 1QRA), the binding pocket of the guanine base in the p21-ras protein was thoroughly examined to identify all of the modes of the non-bonded interactions, including hydrogen bonding, salt bridge interactions, π–π stacking interactions, cation–π interactions, CH–π interactions, and XH–π interactions (XH = NH, OH, and SH).
Figure 7 shows the modes of the non-bonded interactions between the guanine base and its interacting residues in the p21-ras protein (PDB ID: 1QRA). The guanine base interacts with its target protein p21-ras via hydrogen bonding, π–π stacking interactions, and cation–π interactions. Either the main chain or the side chain of a residue can form hydrogen bond with guanine. As shown in Figure 7a, there exist multiple hydrogen bonds between the guanine base and the side chains of the Asn116 and Asp119 residues. Interestingly, the carboxyl group of Asp119 forms dual-hydrogen bonds with guanine, where the N2 and N1 atoms of guanine acts as a hydrogen bond donors. In addition, the main chain amino groups of Lys117 and Ala146 donate their hydrogen to the O6 of the guanine ring to form multiple hydrogen bond interactions. The aromatic residue Phe28 is well positioned for π–π stacking interactions with the guanine ring. The ε-amino groups of positively charged Lys117 and Lys147 are involved in cation–π interactions with guanine.
Subsequently, the strengths of the non-bonded interaction energies between guanine and its surrounding protein residues were quantified in a pairwise manner by means of the double-hybrid DFT method B2PLYP-D3/cc-pVDZ (see Theory and Methods for details). The resulting pairwise intermolecular interaction energies between guanine and surrounding residues are listed in Table 5. As shown in the table, the most significant contributor to the interaction energy ( Δ E I n t a q ) for guanine binding comes from cation–π interactions involving residue Lys117 and Lys147. These interactions account for −18.2 kcal/mol, representing 59.6% of the total binding energy. It is worth noting here that, in addition to the cation–π interactions originating from the positively charged ε-amino groups of lysines, the side-chain alkyl groups of lysine that are parallel to the guanine ring can form multiple CH–π interactions [38] with the guanine ring. The latter enhances the overall strength of the non-bonded interactions involving the lysine residues, as suggested in Ref. [39]. Interestingly, Lys117 also contributes to binding via hydrogen bonds from its main-chain amino group. Hydrogen bonds contribute a total of −9.3 kcal/mol to guanine binding, representing 30.5% of the overall binding energy. This contribution originates from the Asn116 (side chain), Lys117 (main chain), Asp119 (side chain), and Ala146 (main chain) residues. As shown in Table 5, the strengths of those hydrogen bonds vary widely, depending on both the distance and angle. The hydrogen bond energy between Asp119 and guanine was found to be the highest among the interactions analyzed, and it was attributed to its dual interaction mode: both the N1 and N2 groups of guanine donate hydrogen atoms to the oxygen atoms of the Asp119 carboxyl group. In contrast, the hydrogen bond energy between Asn116 and guanine at the N7 position was found to be the weakest. This is due to the extended N–N distance and a suboptimal hydrogen bond angle of 134°, which deviates from linearity and reduces bond strength. Notably, Asn116, Lys117, and Asp119 are part of the NKXD sequence motif (G4 motif), while Ala146 and Lys147 belong to the G5 sequence motif. The π–π stacking interactions between Phe28 and the guanine ring yield an interaction energy of −2.99 kcal/mol.
In summary, the above analysis revealed the energetic hierarchy of the non-bonded interactions in guanine recognition by GTP-binding proteins. Cation–π interactions emerged as the primary source of binding strength, followed by hydrogen bonding for specificity and π–π stacking as an additional stabilizing factor. In particular, the hydrogen bonding interactions between guanine and the side chains of the Asn116 and Asp119 residues, as well as the cation–π interactions between guanine and the positively charged side chains of Lys117 and Lys147, were found to be responsible for the needed specificity and affinity for molecular recognition. In addition, the participation of the main chains of the Lys117 and Ala146 residues in hydrogen bonding interactions with guanine further enhances binding affinity. Furthermore, π–π stacking interactions also meaningfully contribute to guanine binding. These findings are significant as the residues involved in these interactions are derived from the classical NKXD sequence motif and the (T/G)(C/S)A sequence motif (G5 motif). The NKXD motif has evolved as a highly effective binding framework that enables proteins to distinguish guanine from other nucleotides like adenine with remarkable precision. Its specific interactions with guanine’s unique functional groups, combined with a flexible structural arrangement, allow NKXD to achieve high specificity while also supporting diverse binding configurations across protein families.

2.5. Biological Significance

The core principle of molecular recognition is the complementarity between a ligand and its receptor, akin to the “lock and key” model, where the receptor serves as the lock and the ligand acts as the key that forms a specific ligand–receptor complex. Over the years, this lock and key model for the molecular recognition of GTP has been explored across various levels of protein structural hierarchy, including sequence motifs, folds, structural motifs, and intermolecular protein–ligand interactions. In 1987, Dever et al. investigated the structural features that define the GTP-binding domain across nine functionally diverse protein families based on primary sequences [40]. It led to the identification of three consensus sequence motives essential for GTP binding: GXXXXGK, DXXG, and NKXD. These elements are spaced 40–80 amino acids apart in most GTP-binding proteins, aiding in recognizing and binding GTP. Since these consensus sequences are conserved among functionally distinct proteins, including elongation factors, the ras protein family, and G proteins, that work suggests a potential application of these motives to screen GTP-binding function from the primary protein sequences of the unknown protein [40]. The subsequent X-ray crystallographical structural determination of the three-dimensional structures of GTP-binding proteins and their complexes confirmed the structural role of the NKXD motif in guanine binding [7,9,41]. Since then, the NKXD motif has been widely viewed as a fingerprint for GTP-binding proteins [9]. In this study, we conducted an analysis of the molecular recognition of the guanine moiety of GTP in the GTP-binding proteins at the level of non-bonded intermolecular interactions. Traditionally, the NKXD sequence is understood as a key motif for guanine binding, where conserved residues directly participate in hydrogen bonding with guanine [9,35,42]. However, this study identifies a surprising level of variability in the hydrogen bonding roles of NKXD residues. Only 56.7% of complexes containing guanine use the NKXD motif for hydrogen bonding, suggesting a broader structural flexibility than previously recognized. Furthermore, 43.3% of guanine-binding proteins lack the NKXD motif entirely, yet they still achieve guanine recognition through alternative hydrogen bonding arrangements. Specifically, proteins without NKXD often utilize aspartate or glutamate in what is described as the “Di/Ei plus” pattern, while others form non-specific hydrogen bonds with a variety of residues. This expanded understanding of the NKXD motif’s variability and the presence of alternative bonding patterns underscores a structural adaptability in guanine recognition, allowing a wider range of proteins to effectively bind guanine despite lacking the classic NKXD sequence. Thus, from the point of view of molecular recognition, this work strongly supports the widely accepted view that non-bonded interactions are the underlying force behind the molecular recognition of a ligand within a protein. Proteins with entirely different folds can adopt analogous recognition schemes characterized by shared protein–ligand interactions.
The hydrogen bonding characteristics of the guanine base are critical for understanding how proteins differentiate between GTP and ATP, and thus merit further discussion. The analysis above indicates that the N2 and O6 atoms in guanine are among the most commonly involved in hydrogen bonding. These specific hydrogen bonds not only stabilize guanine, but also prevent similar binding with adenine as it lacks the N2 and O6 atoms. In contrast, ATP-binding proteins predominantly engage the N1 and N6 atoms of the adenine base for hydrogen bonding [24]. Bear in mind that the N1 atom of guanine is a hydrogen bond donor while that of adenine is an acceptor. Thus, the NKXD motif confers a selective advantage, allowing proteins to differentiate guanine with high fidelity, which is essential for processes where precise nucleotide recognition underpins cellular signaling and function.
We deciphered the molecular determinants involved in the intermolecular recognition of the guanine moiety of GTP by proteins. Our focus is on understanding the types of interactions employed by enzymes for the recognition of the guanine base and their relative importance. In addition to confirming the importance of well-established hydrogen bonding, we found that two additional forms of non-bonded interactions—π–π stacking and cation–π interactions—are also crucial for the guanine binding in proteins. High-level density functional theory (DFT) calculations further support this by demonstrating the significant contributions of hydrogen bonding, π–π stacking, and cation–π interactions to the overall binding affinity of GTP within proteins. It is important to note that previous studies on protein–ligand interactions, particularly those involving GTP, have primarily emphasized hydrogen bonding and hydrophobic interactions [9,40]. However, the data mining and quantum chemical analyses presented here clearly indicate that π–π stacking and cation–π interactions also play critical roles in the binding of the guanine moiety of GTP to proteins.

3. Theory and Methods

3.1. Data Mining

To establish a database of GTP-binding proteins, a comprehensive data mining of the Protein Data Bank was performed (https://www.rcsb.org). We focused on high-resolution crystal structures (2.5 Å or better) and excluded proteins with over 90% sequence identity to minimize redundancy. Only structures bound to GTP, GDP, or GMP were considered. This resulted in 298 distinct high-resolution crystal structures of GTP-binding protein complexes.

3.2. Analysis of Interaction Modes

First, the crystal structures of all 298 GTP-binding protein complexes were aligned by the superimposition of the guanine base using the Visual Molecular Dynamics (VMD) program [43]. Then, the non-bonded interaction modes, i.e., hydrogen bond, cation–π interaction, and π–π stacking interactions between each guanine base and its surrounding residues in each of the 298 complexes, were systematically analyzed to decipher the specific interactions responsible for molecular recognition. A database of such interaction modes was established, with complete details listed in Table S1.

3.3. Quantification of Intermolecular Interaction Energy

The framework for the ligand–protein complex formation in solution is illustrated by the following scheme:
P ( aq )                 +                 L ( aq )         Δ E i n t   a q         P L ( a q ) Δ G P s o l                           Δ G L s o l                                 Δ G P L s o l . P ( g )                 +                 L ( g )                   Δ E i n t   g   P L ( g )
This scheme underpins our analysis of guanine–protein binding affinities. Similar schemes have been used for the solution-phase binding affinity calculations of ligand–protein complexes in previous works [20,22].
Proteins and ligands lose part of their solvation shell upon binding, incurring dehydration energy. The binding energy in solution is thus evaluated via gas-phase intermolecular interaction energies Δ E I n t g a s corrected for dehydration energy Δ E D e h :
Δ E I n t a q = Δ E I n t g a s + Δ E D e h
Gas-phase interaction energies were calculated using the supermolecular approach. In the supermolecular approach, the gas-phase energy of the interaction between molecules P and L is defined as the difference between the energy of the interacting dimer E P L and the sum of the energies of monomers E P and E L .
Δ E I n t g a s = E P L ( E P + E L ) .
The intermolecular interaction energy calculations were performed using Gaussian 09 software by means of the B2PLYP double-hybrid functional [31,44] with Grimme’s D3BJ dispersion correction [32] in conjunction with the cc-pVDZ basis set [33] (B2PLYP-D3/cc-pVDZ). The basis set superimposition error was corrected by the Boys and Bernardi Counter Poise Method [45].
Dehydration energy is defined as the difference of free energy of solvation:
Δ E D e h = Δ G P L S o l Δ G P S o l Δ G L S o l
Due to high costs of explicit solvent simulations, the free energy of solvation was computed by applying the SM5.42R solvation continuum model by Cramer and Truhlar [46], as implemented in GAMESS [47].

4. Conclusions

In this study, we deciphered the molecular determinants essential for guanine recognition in GTP-binding proteins using a multifaceted approach, encompassing large-scale data mining, in-depth analysis of interaction modes, and rigorous quantum chemical calculations. It was found that multiple modes of non-bonded interactions are employed by GTP-binding proteins to achieve molecular recognition. Hydrogen bonds lock guanine in place with specificity, while cation–π interactions provide strong electrostatic interaction support, and π–π stacking further stabilizes the binding complex.
  • Hydrogen bonds, particularly those involving N2 and O6 atoms of the guanine base, confer specificity to guanine recognition by distinguishing it from adenine.
  • Quantum chemical analysis revealed the critical role of cation–π interactions between the guanine ring and its surrounding basic residues (Lys and Arg) in stabilizing guanine binding within GTP-binding proteins. Intermolecular interaction energies for representative cation–π interactions range from −1.51 to −10.61 kcal/mol. The high-energy strength of cation–π interactions can be attributed to the multi-mode intermolecular interactions associated with the Lys and Arg residues. For example, the Lys residue of the NKXD motif can be involved in both the cation–π interactions between the positively charged ε-amino groups of lysine and the guanine ring, as well as in the CH–π interactions between the side chain alkyl groups of lysine and the guanine ring.
  • π–π stacking interactions between the guanine ring and its surrounding aromatic residues (Phe, Tyr, and Trp) act as an auxiliary stabilizing factor. In complex featuring the NKXD motif, those aromatic residues are typically situated on the opposite side of the guanine ring relative to the Lys residue of the NKXD motif (see, for example, the table of content figure). Intermolecular interaction energies for representative π–π stacking interactions range from −0.34 to −6.57 kcal/mol.
This combination of non-bonded interaction modes maximizes both the strength and selectivity of the molecular recognition of guanine in GTP-binding proteins.
The collective insights gleaned from these investigations illuminate the sophisticated molecular recognition strategies employed by GTP-binding proteins. By harnessing a combination of hydrogen bonding, cation–π, and π–π stacking interactions, these proteins achieve the requisite specificity and affinity for effective guanine binding. This versatile interaction framework not only stabilizes guanine within diverse protein families, but also underpins the essential biological functions of GTP-binding proteins in various cellular processes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms252212449/s1.

Author Contributions

Conceptualization, X.H.; methodology, X.H. and P.B.; software, X.H. and P.B.; validation, P.B.; formal analysis, P.B.; investigation, P.B.; resources, X.H.; data curation, P.B.; writing—original draft preparation, P.B.; writing—review and editing, X.H.; visualization, P.B.; supervision, X.H.; project administration, X.H.; funding acquisition, X.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the relevant data are provided in Table 1 of the main text and Table S1 of the Supplementary Materials.

Acknowledgments

We are pleased to acknowledge the Ohio Supercomputer Center for a generous allocation of supercomputer time.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Pfeuffer, T.; Helmreich, E.J.M. Structural and Functional-Relationships of Guanosine Triphosphate Binding-Proteins. Curr. Top. Cell. Regul. 1988, 29, 129–216. [Google Scholar] [PubMed]
  2. Spiegel, A.M. Signal Transduction by Guanine-Nucleotide Binding-Proteins. Mol. Cell. Endocrinol. 1987, 49, 1–16. [Google Scholar] [CrossRef] [PubMed]
  3. Bockaert, J.; Homburger, V.; Rouot, B. Gtp Binding-Proteins—A Key Role in Cellular Communication. Biochimie 1987, 69, 329–338. [Google Scholar] [CrossRef] [PubMed]
  4. Bourne, H.R.; Sanders, D.A.; McCormick, F. The Gtpase Superfamily—A Conserved Switch for Diverse Cell Functions. Nature 1990, 348, 125–132. [Google Scholar] [CrossRef] [PubMed]
  5. Li, G.; Zhang, X.C. GTP Hydrolysis Mechanism of Ras-like GTPases. J. Mol. Biol. 2004, 340, 921–932. [Google Scholar] [CrossRef]
  6. Guidolin, D.; Agnati, L.F.; Marcoli, M.; Borroto-Escuela, D.O.; Fuxe, K. G-protein-coupled receptor type A heteromers as an emerging therapeutic target. Expert Opin. Ther. Targets 2015, 19, 265–283. [Google Scholar] [CrossRef]
  7. Bourne, H.R.; Sanders, D.A.; McCormick, F. The Gtpase Superfamily—Conserved Structure And Molecular Mechanism. Nature 1991, 349, 117–127. [Google Scholar] [CrossRef]
  8. Sprang, S.R. G protein mechanisms: Insights from structural analysis. Annu. Rev. Biochem. 1997, 66, 639–678. [Google Scholar] [CrossRef]
  9. Kjeldgaard, M.; Nyborg, J.; Clark, B.F. The GTP binding motif: Variations on a theme. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 1996, 10, 1347–1368. [Google Scholar] [CrossRef]
  10. Babor, M.; Sobolev, V.; Edelman, M. Conserved positions for ribose recognition: Importance of water bridging interactions among ATP, ADP and FAD-protein complexes. J. Mol. Biol. 2002, 323, 523–532. [Google Scholar] [CrossRef]
  11. Dreusicke, D.; Schulz, G.E. The glycine-rich loop of adenylate kinase forms a giant anion hole. FEBS Lett. 1986, 208, 301–304. [Google Scholar] [CrossRef] [PubMed]
  12. Sprang, S.R. Invited review: Activation of G proteins by GTP and the mechanism of Gα-catalyzed GTP hydrolysis. Biopolymers 2016, 105, 449–462. [Google Scholar] [CrossRef] [PubMed]
  13. Saraste, M.; Sibbald, P.R.; Wittinghofer, A. The P-loop--a common motif in ATP- and GTP-binding proteins. Trends Biochem. Sci. 1990, 15, 430–434. [Google Scholar] [CrossRef] [PubMed]
  14. Hirsch, A.K.; Fischer, F.R.; Diederich, F. Phosphate recognition in structural biology. Angew. Chem. (Int. Ed. Engl.) 2007, 46, 338–352. [Google Scholar] [CrossRef]
  15. Babine, R.E.; Bender, S.L. Molecular recognition of protein-ligand complexes: Applications to drug design. Chem. Rev. 1997, 97, 1359–1472. [Google Scholar] [CrossRef]
  16. McGaughey, G.B.; Gagné, M.; Rappé, A.K. π-Stacking interactions. Alive and well in proteins. J. Biol. Chem. 1998, 273, 15458–15463. [Google Scholar] [CrossRef]
  17. Bohm, H.J.; Klebe, G. What can we learn from molecular recognition in protein-ligand complexes for the design of new drugs? Angew. Chem.-Int. Ed. 1996, 35, 2589–2614. [Google Scholar] [CrossRef]
  18. Neel, A.J.; Hilton, M.J.; Sigman, M.S.; Toste, F.D. Exploiting non-covalent π interactions for catalyst design. Nature 2017, 543, 637–646. [Google Scholar] [CrossRef]
  19. Persch, E.; Dumele, O.; Diederich, F. Molecular Recognition in Chemical and Biological Systems. Angew. Chem. Int. Ed. 2015, 54, 3290–3327. [Google Scholar] [CrossRef]
  20. Zhu, Y.; Alqahtani, S.; Hu, X.C. Aromatic Rings as Molecular Determinants for the Molecular Recognition of Protein Kinase Inhibitors. Molecules 2021, 26, 1776. [Google Scholar] [CrossRef]
  21. Carter-Fenk, K.; Liu, M.L.; Pujal, L.; Loipersberger, M.; Tsanai, M.; Vernon, R.M.; Forman-Kay, J.D.; Head-Gordon, M.; Heidar-Zadeh, F.; Head-Gordon, T. The Energetic Origins of Pi-Pi Contacts in Proteins. J. Am. Chem. Soc. 2023, 145, 24836–24851. [Google Scholar] [CrossRef] [PubMed]
  22. Zhu, Y.; Hu, X.C. Molecular Recognition of FDA-Approved Small Molecule Protein Kinase Drugs in Protein Kinases. Molecules 2022, 27, 7124. [Google Scholar] [CrossRef] [PubMed]
  23. Gallivan, J.P.; Dougherty, D.A. Cation-π interactions in structural biology. Proc. Natl. Acad. Sci. USA 1999, 96, 9459–9464. [Google Scholar] [CrossRef] [PubMed]
  24. Mao, L.; Wang, Y.; Liu, Y.; Hu, X. Molecular determinants for ATP-binding in proteins: A data mining and quantum chemical analysis. J. Mol. Biol. 2004, 336, 787–807. [Google Scholar] [CrossRef]
  25. Dougherty, D.A. The cation-π interaction. Acc. Chem. Res. 2013, 46, 885–893. [Google Scholar] [CrossRef]
  26. Anderson, M.A.; Ogbay, B.; Arimoto, R.; Sha, W.; Kisselev, O.G.; Cistola, D.P.; Marshall, G.R. Relative Strength of Cation-π vs. Salt-Bridge Interactions:  The Gtα(340−350) Peptide/Rhodopsin System. J. Am. Chem. Soc. 2006, 128, 7531–7541. [Google Scholar] [CrossRef]
  27. Kumar, K.; Woo, S.M.; Siu, T.; Cortopassi, W.A.; Duarte, F.; Paton, R.S. Cation–π interactions in protein–ligand binding: Theory and data-mining reveal different roles for lysine and arginine. Chem. Sci. 2018, 9, 2655–2665. [Google Scholar] [CrossRef]
  28. Sivasakthi, V.; Anitha, P.; Kumar, K.M.; Bag, S.; Senthilvel, P.; Lavanya, P.; Swetha, R.; Anbarasu, A.; Ramaiah, S. Aromatic-aromatic interactions: Analysis of π-π interactions in interleukins and TNF proteins. Bioinformation 2013, 9, 432–439. [Google Scholar] [CrossRef]
  29. Deng, J.H.; Luo, J.; Mao, Y.L.; Lai, S.; Gong, Y.N.; Zhong, D.C.; Lu, T.B. π-π stacking interactions: Non-negligible forces for stabilizing porous supramolecular frameworks. Sci. Adv. 2020, 6, eaax9976. [Google Scholar] [CrossRef]
  30. Zhu, Y.; Alqahtani, S.; Hu, X.C. An Assessment of Dispersion-Corrected DFT Methods for Modeling Nonbonded Interactions in Protein Kinase Inhibitor Complexes. Molecules 2024, 29, 304. [Google Scholar] [CrossRef]
  31. Grimme, S.; Neese, F. Double-hybrid density functional theory for excited electronic states of molecules. J. Chem. Phys. 2007, 127, 154116. [Google Scholar] [CrossRef]
  32. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef] [PubMed]
  33. Dunning, T.H. Gaussian-Basis Sets for Use in Correlated Molecular Calculations. 1. The Atoms Boron Through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  34. Mistry, J.; Chuguransky, S.; Williams, L.; Qureshi, M.; Salazar, G.A.; Sonnhammer, E.L.L.; Tosatto, S.C.E.; Paladin, L.; Raj, S.; Richardson, L.J.; et al. Pfam: The protein families database in 2021. Nucleic Acids Res. 2021, 49, D412–D419. [Google Scholar] [CrossRef] [PubMed]
  35. Leipe, D.D.; Wolf, Y.I.; Koonin, E.V.; Aravind, L. Classification and evolution of P-loop GTPases and related ATPases. J. Mol. Biol. 2002, 317, 41–72. [Google Scholar] [CrossRef] [PubMed]
  36. Tsuzuki, S.; Honda, K.; Uchimaru, T.; Mikami, M.; Tanabe, K. Origin of attraction and directionality of the π/π interaction: Model chemistry calculations of benzene dimer interaction. J. Am. Chem. Soc. 2002, 124, 104–112. [Google Scholar] [CrossRef] [PubMed]
  37. Scheidig, A.J.; Burmester, C.; Goody, R.S. The pre-hydrolysis state of p21ras in complex with GTP: New insights into the role of water molecules in the GTP hydrolysis reaction of ras-like proteins. Structure 1999, 7, 1311–1324. [Google Scholar] [CrossRef]
  38. Xiao, Y.; Woods, R.J. Protein-Ligand CH-π Interactions: Structural Informatics, Energy Function Development, and Docking Implementation. J. Chem. Theory Comput. 2023, 19, 5503–5515. [Google Scholar] [CrossRef]
  39. Mao, L.S.; Wang, Y.L.; Liu, Y.M.; Hu, X.C. Multiple intermolecular interaction modes of positively charged residues with adenine in ATP-binding proteins. J. Am. Chem. Soc. 2003, 125, 14216–14217. [Google Scholar] [CrossRef]
  40. Dever, T.E.; Glynias, M.J.; Merrick, W.C. GTP-Binding Domain-3 Consensus Sequence Elements with Distinct Spacing. Proc. Natl. Acad. Sci. USA 1987, 84, 1814–1818. [Google Scholar] [CrossRef]
  41. Takai, Y.; Sasaki, T.; Matozaki, T. Small GTP-binding proteins. Physiol. Rev. 2001, 81, 153–208. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, S.; Cerione, R.A.; Clardy, J. Structural basis for the guanine nucleotide-binding activity of tissue transglutaminase and its regulation of transamidation activity. Proc. Natl. Acad. Sci. USA 2002, 99, 2743–2747. [Google Scholar] [CrossRef]
  43. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef] [PubMed]
  44. Grimme, S. Semiempirical hybrid density functional with perturbative second-order correlation. J. Chem. Phys. 2006, 124, 034108. [Google Scholar] [CrossRef] [PubMed]
  45. Boys, S.F.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  46. Li, J.; Zhu, T.; Hawkins, G.D.; Winget, P.; Liotard, D.A.; Cramer, C.J.; Truhlar, D.G. Extension of the platform of applicability of the SM5. 42R universal solvation model. Theor. Chem. Acc. 1999, 103, 9–63. [Google Scholar] [CrossRef]
  47. Schmidt, M.W.; Baldridge, K.K.; Boatz, J.A.; Elbert, S.T.; Gordon, M.S.; Jensen, J.H.; Koseki, S.; Matsunaga, N.; Nguyen, K.A.; Su, S. General atomic and molecular electronic structure system. J. Comput. Chem. 1993, 14, 1347–1363. [Google Scholar] [CrossRef]
Figure 1. (a) The guanine base of the guanine nucleotide (GTP/GDP/GMP), where the symbol R represents ribose and phosphate groups. The inward arrow shows the hydrogen bond acceptor, and the outward arrow shows the hydrogen bond donor. All the atoms are labeled according to the IUPAC naming system. (b) Structure of a representative GTP-binding protein: a p21-ras protein bound to GTP (PDB ID:1QRA).
Figure 1. (a) The guanine base of the guanine nucleotide (GTP/GDP/GMP), where the symbol R represents ribose and phosphate groups. The inward arrow shows the hydrogen bond acceptor, and the outward arrow shows the hydrogen bond donor. All the atoms are labeled according to the IUPAC naming system. (b) Structure of a representative GTP-binding protein: a p21-ras protein bound to GTP (PDB ID:1QRA).
Ijms 25 12449 g001
Figure 2. Representative hydrogen bond patterns: (a) Ni-Ki+1-Xi+2-Di+3 in the p21-ras protein (PDB ID: 1QRA); (b) Ni-Ki+1-Xi+2-Di+3 in the Human Ras-like, family 12 protein (PDB ID: 3C5C); (c) Ni-Ki+1-Xi+2-Di+3 in the human adenylosuccinate synthetase isozyme 2 (PDB ID: 2V40); (d) Ni-Ki+1-Xi+2-Di+3 in the Plasmodium falciparum rab6 protein (PDB ID: 1D5C); (e) “Di/Ei plus” in PnrA from Treponema pallidum (PDB ID: 2FQX); and (f) the “Others” pattern in Murray Valley encephalitis virus methyltransferase domain (PDB ID: 2PXA). C, N, O, and S atoms are colored in cyan, blue, red and yellow, respectively.
Figure 2. Representative hydrogen bond patterns: (a) Ni-Ki+1-Xi+2-Di+3 in the p21-ras protein (PDB ID: 1QRA); (b) Ni-Ki+1-Xi+2-Di+3 in the Human Ras-like, family 12 protein (PDB ID: 3C5C); (c) Ni-Ki+1-Xi+2-Di+3 in the human adenylosuccinate synthetase isozyme 2 (PDB ID: 2V40); (d) Ni-Ki+1-Xi+2-Di+3 in the Plasmodium falciparum rab6 protein (PDB ID: 1D5C); (e) “Di/Ei plus” in PnrA from Treponema pallidum (PDB ID: 2FQX); and (f) the “Others” pattern in Murray Valley encephalitis virus methyltransferase domain (PDB ID: 2PXA). C, N, O, and S atoms are colored in cyan, blue, red and yellow, respectively.
Ijms 25 12449 g002
Figure 3. A 3D stereographic drawing of a guanine base surrounded by positively charged residues. All the 258 complexes that contain cation–π interactions are aligned by the superimposition of the guanine base.
Figure 3. A 3D stereographic drawing of a guanine base surrounded by positively charged residues. All the 258 complexes that contain cation–π interactions are aligned by the superimposition of the guanine base.
Ijms 25 12449 g003
Figure 4. Representative cation–π interactions between the positively charged residues and the guanine ring. PDB IDs for the cation–π interacting motifs are displayed. C, N, and O atoms are colored in cyan, blue, and red, respectively. The red dashed line indicates the coexistence of hydrogen bond interactions.
Figure 4. Representative cation–π interactions between the positively charged residues and the guanine ring. PDB IDs for the cation–π interacting motifs are displayed. C, N, and O atoms are colored in cyan, blue, and red, respectively. The red dashed line indicates the coexistence of hydrogen bond interactions.
Ijms 25 12449 g004
Figure 5. A 3D stereographic drawing of a guanine base surrounded by aromatic residues. All of the 162 complexes that contain π–π stacking interactions are aligned by the superimposition of the guanine base.
Figure 5. A 3D stereographic drawing of a guanine base surrounded by aromatic residues. All of the 162 complexes that contain π–π stacking interactions are aligned by the superimposition of the guanine base.
Ijms 25 12449 g005
Figure 6. Representative π–π stacking interactions between the aromatic residue and the guanine ring. PDB IDs for the π–π stacking interacting motifs are displayed. C, N, and O atoms are colored in cyan, blue, and red, respectively.
Figure 6. Representative π–π stacking interactions between the aromatic residue and the guanine ring. PDB IDs for the π–π stacking interacting motifs are displayed. C, N, and O atoms are colored in cyan, blue, and red, respectively.
Ijms 25 12449 g006
Figure 7. (a) A schematic intermolecular interaction map between the guanine and its interacting residues in a GTP-binding protein p21-ras (PDB ID: 1QRA). The interatomic distances (in Å) are indicated along the dashed lines. The red, blue, and gray dashed lines represent hydrogen bond interactions, cation–π interactions, and π–π stacking interactions, respectively. (b) The 3D structure of the residues surrounding the guanine. For clarity, only side-chain interactions are shown. The color codes of the dashed lines are the same as in (a).
Figure 7. (a) A schematic intermolecular interaction map between the guanine and its interacting residues in a GTP-binding protein p21-ras (PDB ID: 1QRA). The interatomic distances (in Å) are indicated along the dashed lines. The red, blue, and gray dashed lines represent hydrogen bond interactions, cation–π interactions, and π–π stacking interactions, respectively. (b) The 3D structure of the residues surrounding the guanine. For clarity, only side-chain interactions are shown. The color codes of the dashed lines are the same as in (a).
Ijms 25 12449 g007
Table 1. List of protein complexes with bound guanine nucleotides.
Table 1. List of protein complexes with bound guanine nucleotides.
PDB IDProteinResolution
(Å)
Nucleotide aFamily b
5DN8GTPase Der1.76GDP50S ribosome-binding GTPase
5X4BGTPase Der1.50GDP50S ribosome-binding GTPase
5M04GTPase ObgE/CgtA1.85GDP50S ribosome-binding GTPase
2DBYGTP-binding protein1.76GDP50S ribosome-binding GTPase
2DYKGTP-binding protein1.96GDP50S ribosome-binding GTPase
4DCUGTP-binding protein ENGA2.00GDP50S ribosome-binding GTPase
3EC1GTP-binding protein YqeH2.36GDP50S ribosome-binding GTPase
1SVIGTP-binding protein YSXC1.95GDP50S ribosome-binding GTPase
1MKYProbable GTP-binding protein EngA1.90GDP50S ribosome-binding GTPase
3PQCProbable GTP-binding protein EngB1.90GDP50S ribosome-binding GTPase
5UCVProbable GTP-binding protein EngB1.80GDP50S ribosome-binding GTPase
3CNOPutative uncharacterized protein2.30GDP50S ribosome-binding GTPase
3QXXDethiobiotin synthetase1.36GDPAAA domain
2FQXMembrane lipoprotein tmpC1.70GMPABC transporter substrate-binding protein PnrA-like
4FWPPropionate kinase2.50GDPAcetokinase family
2QVNAdenosine deaminase2.19GMPAdenosine deaminase
1LONAdenylosuccinate synthetase2.10GDPAdenylosuccinate synthetase
1P9BAdenylosuccinate Synthetase2.00GDPAdenylosuccinate synthetase
1QF5Adenylosuccinate synthetase2.00GDPAdenylosuccinate synthetase
2V40Adenylosuccinate synthetase1.90GDPAdenylosuccinate synthetase
5I34Adenylosuccinate synthetase1.53GDPAdenylosuccinate synthetase
6C25Adenylosuccinate synthetase1.90GDPAdenylosuccinate synthetase
2X77ADP-ribosylation factor2.10GDPADP-ribosylation factor family
1R8SADP-ribosylation factor 11.46GDPADP-ribosylation factor family
3AQ4ADP-ribosylation factor 11.80GDPADP-ribosylation factor family
3LRPADP-ribosylation factor 12.50GDPADP-ribosylation factor family
4Y0VADP-ribosylation factor 11.80GDPADP-ribosylation factor family
2B6HADP-ribosylation factor 51.76GDPADP-ribosylation factor family
2A5DADP-ribosylation factor 61.80GTPADP-ribosylation factor family
1UPTADP-ribosylation factor-like protein 11.70GTPADP-ribosylation factor family
1ZD9ADP-ribosylation factor-like 10B1.70GDPADP-ribosylation factor family
1YZGADP-ribosylation factor-like 82.00GDPADP-ribosylation factor family
2H17ADP-ribosylation factor-like protein 5A1.70GDPADP-ribosylation factor family
2H57ADP-ribosylation factor-like protein 62.00GTPADP-ribosylation factor family
1FZQADP-ribosylation factor-like protein31.70GDPADP-ribosylation factor family
4V0KARF-LIKE SMALL GTPASE1.438GDPADP-ribosylation factor family
3T1OGliding protein mglA1.90GDPADP-ribosylation factor family
5UF8Potential ADP-ribosylation factor1.87GDPADP-ribosylation factor family
3BH7Protein XRP21.90GDPADP-ribosylation factor family
1F6BSAR11.70GDPADP-ribosylation factor family
1H65CHLOROPLAST OUTER ENVELOPE PROTEIN OEP342.00GDPAIG1 family
3V70GTPase IMAP family member 12.21GDPAIG1 family
2XTMGTPASE IMAP FAMILY MEMBER 21.70GDPAIG1 family
3LXXGTPase IMAP family member 42.15GDPAIG1 family
3DEFT7I23.11 protein1.96GDPAIG1 family
4A7WURIDYLATE KINASE1.80GTPAmino acid kinase family
3ZF8MANNAN POLYMERASE COMPLEXES SUBUNIT MNN91.98GDPAnp1
2FP4Succinyl-CoA ligase [GDP-forming] alpha-chain2.08GTPATP-grasp domain
3UFXsuccinyl-CoA synthetase alpha subunit2.35GDPATP-grasp domain
4Z87Inosine-5′-monophosphate dehydrogenase2.25GDPCBS domain
2Q0ERNA uridylyl transferase2.10GTPCid1 family poly A polymerase
4LPSHydrogenase/urease nickel incorporation protein HypB2.00GDPCobW/HypB/UreG, nucleotide-binding domain
4HI0Urease accessory protein UreF2.35GDPCobW/HypB/UreG, nucleotide-binding domain
5EY0GTP-sensing transcriptional pleiotropic repressor CodY1.60GTPCodY GAF-like domain
1YRBATP(GTP)binding protein1.75GDPConserved hypothetical ATP binding protein
5HCIGPN-loop GTPase 12.30GDPConserved hypothetical ATP binding protein
2CXXProbable GTP-binding protein engB1.70GDPC-terminal region of MMR_HSR1 domain
3LZZPutative uncharacterized protein2.50GDPCupin superfamily (DUF985)
1IH7DNA primase2.21GMPDNA polymerase family B
4EDKDNA primase2.00GTPDNA primase catalytic core, N-terminal domain
4B2PDNA recombination and repair protein Rad51-like1.60GTPDNA recombination/repair protein RadA
5D3QDynamin-1,Dynamin-11.70GDPDynamin family
3W6PDynamin-1-like protein1.70GDPDynamin family
3L43Dynamin-32.27GDPDynamin family
4P4TInterferon-induced GTP-binding protein Mx12.30GDPDynamin family
1KK3eIF2gamma1.90GDPElongation factor Tu GTP-binding domain
4TMXeIF5B1.50GTPElongation factor Tu GTP-binding domain
1IJEElongation factor 1-alpha2.40GDPElongation factor Tu GTP-binding domain
1SKQElongation factor 1-alpha1.80GDPElongation factor Tu GTP-binding domain
3WXMElongation factor 1-alpha2.30GTPElongation factor Tu GTP-binding domain
3WY9Elongation factor 1-alpha2.30GDPElongation factor Tu GTP-binding domain
5H7KElongation factor 21.60GDPElongation factor Tu GTP-binding domain
2BM0Elongation factor G2.40GDPElongation factor Tu GTP-binding domain
2DY1Elongation factor G1.60GTPElongation factor Tu GTP-binding domain
1D2EElongation factor Tu1.94GDPElongation factor Tu GTP-binding domain
1HA3Elongation factor Tu2.00GDPElongation factor Tu GTP-binding domain
4J0QElongation factor Tu-A2.30GDPElongation factor Tu GTP-binding domain
4NCNEukaryotic translation initiation factor 5B-like protein1.87GTPElongation factor Tu GTP-binding domain
2YWHGTP-binding protein LepA2.24GDPElongation factor Tu GTP-binding domain
3TR5Peptide chain release factor 32.11GDPElongation factor Tu GTP-binding domain
3VQTPeptide chain release factor 31.80GDPElongation factor Tu GTP-binding domain
5FG3Probable translation initiation factor IF-21.90GDPElongation factor Tu GTP-binding domain
2HCJProtein chain elongation factor EF-Tu2.12GDPElongation factor Tu GTP-binding domain
4ZKDSuperkiller protein 72.18GDPElongation factor Tu GTP-binding domain
4B47Translation initiation factor IF-22.30GDPElongation factor Tu GTP-binding domain
1G7STranslation initiation factor IF2/EIF5B2.00GDPElongation factor Tu GTP-binding domain
4RD1Translation initiation factor 2 subunit gamma1.50GDPElongation factor Tu GTP-binding domain
2PHNF420-0:gamma-glutamyl ligase1.35GDPF420-0:Gamma-glutamyl ligase
3I8XFerrous iron transport protein B2.25GDPFerrous iron transport protein B
3W5JFerrous iron transport protein B1.93GDPFerrous iron transport protein B
2WJHFerrous iron transport protein B HOMOLOG2.10GDPFerrous iron transport protein B
4NONFerrous iron uptake transporter protein B2.50GDPFerrous iron transport protein B
3A1SIron(II) transport protein B1.50GDPFerrous iron transport protein B
4HDGPolyprotein2.38GTPFlavivirus RNA-directed RNA polymerase
5VYRFormyltransferase1.70GMPFormyl transferase
2PXAGenome polyprotein2.30GTPFtsJ-like methyltransferase
4V0RNS5 POLYMERASE2.40GTPFtsJ-like methyltransferase
3EVDRNA-directed RNA polymerase NS51.50GTPFtsJ-like methyltransferase
5GOZRNA-directed RNA polymerase NS52.05GTPFtsJ-like methyltransferase
5U32tRNA ligase2.19GDPFungal tRNA ligase phosphodiesterase domain
3ZY2GDP-fucose protein O-fucosyltransferase 11.54GDPGDP-fucose protein O-fucosyltransferase
5KXHGDP-fucose protein O-fucosyltransferase 11.33GDPGDP-fucose protein O-fucosyltransferase
5FOEGDP-fucose protein O-fucosyltransferase 2,Thrombospondin-11.98GDPGDP-fucose protein O-fucosyltransferase
2Z1MGDP-D-mannose dehydratase2.00GDPGDP-mannose 4,6 dehydratase
1N7HGDP-D-mannose-4,6-dehydratase1.80GDPGDP-mannose 4,6 dehydratase
5IN4GDP-mannose 4,6 dehydratase1.60GDPGDP-mannose 4,6 dehydratase
1RPNGDP-mannose 4,6-dehydratase2.15GDPGDP-mannose 4,6 dehydratase
5UZHNafoA.00085.b2.25GDPGDP-mannose 4,6 dehydratase
6DHMGlutamate dehydrogenase 1, mitochondrial3.00GTPGlutamate/Leucine/Phenylalanine/Valine dehydrogenase
1S4OGlycolipid 2-alpha-mannosyltransferase2.01GDPGlycolipid 2-alpha-mannosyltransferase
5A07Mannosyltransferase KTR41.90GDPGlycolipid 2-alpha-mannosyltransferase
5MLZDolichol monophosphate mannose synthase2.00GDPGlycosyl transferase family 2
2Y4KMANNOSYLGLYCERATE SYNTHASE2.45GDPGlycosyl transferase family 2
3OKC Mannosyltransferase2.40GDPGlycosyl transferases group 1
4N9W Mannosyltransferase1.94GDPGlycosyl transferases group 1
2NZXAlpha1,3-fucosyltransferase1.90GDPGlycosyltransferase family 10 (fucosyltransferase) C-term
4F97VldE2.11GDPGlycosyltransferase family 20
1ZCBG alpha i/132.00GDPG-protein alpha subunit
2ODEGuanine nucleotide-binding protein G(k) subunit alpha1.90GDPG-protein alpha subunit
1TADTRANSDUCIN-ALPHA1.70GDPG-protein alpha subunit
4DU6GTP cyclohydrolase 12.11GTPGTP cyclohydrolase I
1A8RGTP CYCLOHYDROLASE I2.11GTPGTP cyclohydrolase I
2QV6GTP cyclohydrolase III2.00GTPGTP cyclohydrolase III
2QTHGTP-binding protein2.00GDPGTP-binding GTPase Middle Region
1ZNYGuanylate kinase2.30GDPGuanylate kinase
2AN9Guanylate kinase2.35GDPGuanylate kinase
6B9FAtlastin-11.90GDPGuanylate-binding protein, N-terminal domain
5VGRAtlastin-32.10GDPGuanylate-binding protein, N-terminal domain
1VJ7Bifunctional RELA/SPOT2.10GDPHD domain
4TNPDeoxynucleoside triphosphate triphosphohydrolase SAMHD12.00GTPHD domain
2HEKHypothetical protein1.99GDPHD domain
2OGIHypothetical protein SAG16611.85GDPHD domain
4TZ0ATP-dependent RNA helicase MSS116, mitochondrial2.35GDPHelicase conserved C-terminal domain
4XBAAprataxin-like protein1.50GMPHIT domain
5AQKHEAT SHOCK COGNATE 71 KDA PROTEIN2.09GDPHsp70 protein
4Q46Polymerase basic protein 21.80GDPInfluenza RNA-dependent RNA polymerase subunit PB2
4LV5Rhoptry protein 5B1.70GDPInterferon-inducible GTPase (IIGP)
5GMFToll-like receptor 72.50GMPLeucine-rich repeat
4NXVMitochondrial dynamic protein MID512.30GDPMab-21 protein
2ZU9Mannosyl-3-phosphoglycerate synthase2.00GDPMannosyl-3-phosphoglycerate synthase (osmo_MPGsynth)
3NXSLAO/AO transport system ATPase2.30GDPMethylmalonyl Co-A mutase-associated GTPase MeaB
3TK1Membrane ATPase/protein kinase2.40GDPMethylmalonyl Co-A mutase-associated GTPase MeaB
4LC1Methylmalonyl-CoA mutase accessory protein1.80GDPMethylmalonyl Co-A mutase-associated GTPase MeaB
3P32Probable GTPase Rv1496/MT15431.90GDPMethylmalonyl Co-A mutase-associated GTPase MeaB
3DMHProbable ribosomal RNA small subunit methyltransferase1.55GMPMethyltransferase small domain
1FRWMOLYBDOPTERIN-GUANINE DINUCLEOTIDE BIOSYNTHESIS PROTEIN1.75GTPMobA-like NTP transferase domain
2FB3Molybdenum cofactor biosynthesis protein A2.35GTPMolybdenum Cofactor Synthesis C
1SIWRespiratory nitrate reductase 1 alpha chain2.20GDPMolybdopterin oxidoreductase
1CKMMRNA CAPPING ENZYME2.50GTPmRNA capping enzyme, catalytic domain
4PZ6mRNA-capping enzyme subunit alpha2.41GMPmRNA capping enzyme, catalytic domain
1JWYMyosin-2 heavy chain, Dynamin-A2.30GDPMyosin head (motor domain)
3SIWNodulation fucosyltransferase NodZ1.98GDPNodulation protein Z (NodZ)
2E87Hypothetical protein PH13202.35GDPNOG1 N-terminal helical domain
2DXENucleoside diphosphate kinase1.70GDPNucleoside diphosphate kinase
3BBFNucleoside diphosphate kinase B1.70GDPNucleoside diphosphate kinase
1RYAGDP-mannose mannosyl hydrolase1.30GDPNUDIX domain
2A8SU8 snoRNA-binding protein X292.45GTPNUDIX domain
4LC4Probable sugar kinase protein1.70GMPpfkB family carbohydrate kinase
2FAHPhosphoenolpyruvate carboxykinase2.09GDPPhosphoenolpyruvate carboxykinase C-terminal P-loop domain
4R43Phosphoenolpyruvate carboxykinase [GTP]1.80GDPPhosphoenolpyruvate carboxykinase C-terminal P-loop domain
1JE15′-METHYLTHIOADENOSINE PHOSPHORYLASE1.80GMPPhosphorylase superfamily
1ODJPURINE NUCLEOSIDE PHOSPHORYLASE2.40GMPPhosphorylase superfamily
3IEXPurine-nucleoside phosphorylase2.05GMPPhosphorylase superfamily
4DT9APH(2″)-Id2.10GMPPhosphotransferase enzyme family
4ORKBifunctional AAC/APH2.30GDPPhosphotransferase enzyme family
3TDWGentamicin resistance protein1.70GDPPhosphotransferase enzyme family
5IGIMacrolide 2′-phosphotransferase1.20GMPPhosphotransferase enzyme family
5IH1Macrolide 2′-phosphotransferase II1.31GDPPhosphotransferase enzyme family
5UXCPredicted aminoglycoside phosphotransferase1.72GDPPhosphotransferase enzyme family
3LDUPutative methylase1.70GTPPutative RNA methylase family UPF0020
5JCPArf-GAP with Rho-GAP domain2.10GDPRas family
5OECGtgE2.30GDPRas family
2G77GTPase-activating protein GYP12.26GDPRas family
4DJTGTP-binding nuclear protein GSP11.80GDPRas family
3M1IGTP-binding nuclear protein GSP1/CNR12.00GTPRas family
3GJ0GTP-binding nuclear protein Ran1.48GDPRas family
2GF0GTP-binding protein Di-Ras11.90GDPRas family
2ERXGTP-binding protein Di-Ras21.65GDPRas family
2G3YGTP-binding protein GEM2.40GDPRas family
2DPXGTP-binding protein RAD1.80GDPRas family
2NZJGTP-binding protein REM 12.50GDPRas family
3CBQGTP-binding protein REM 21.82GDPRas family
6BSXGTP-binding protein Rheb1.65GDPRas family
4KLZGTP-binding protein Rit12.30GDPRas family
3RWOGTP-binding protein YPT32/YPT111.70GDPRas family
1KY3GTP-BINDING PROTEIN YPT7P1.35GDPRas family
2ZEJLeucine-rich repeat kinase 22.00GDPRas family
5UB8Likely Rab family GTP-binding protein2.35GDPRas family
2WKQNPH1-1, RAS-RELATED C3 BOTULINUM TOXIN SUBSTRATE 11.60GTPRas family
1QRAP21RAS1.60GTPRas family
5XC5Probable Rab-related GTPase1.40GTPRas family
1EK0PROTEIN (GTP-BINDING PROTEIN YPT51)1.48GDPRas family
2F9LRAB11B, member RAS oncogene family1.55GDPRas family
2IL1Rab122.10GDPRas family
1Z0FRAB14, member RAS oncogene family2.15GDPRas family
3CLVRab5 protein, putative1.89GDPRas family
1D5CRAB6 GTPASE2.30GDPRas family
3BWDRac-like GTP-binding protein ARAC61.53GDPRas family
2J0VRAC-LIKE GTP-BINDING PROTEIN ARAC71.78GDPRas family
1KAORAP2A1.70GDPRas family
2P5SRAS and EF-hand domain containing2.15GDPRas family
2Q3HRas homolog gene family, member U1.73GDPRas family
5WDSRas protein1.85GDPRas family
2ATVRAS-like estrogen-regulated growth inhibitor1.90GDPRas family
3C5CRAS-like protein 121.85GDPRas family
5O33Ras-related C3 botulinum toxin substrate 11.64GDPRas family
5VCURas-related c3 botulinum toxin substrate 1 isoform x21.85GDPRas family
3KKQRas-related protein M-Ras1.20GDPRas family
1Z0IRas-related protein Rab-212.33GDPRas family
1Z0JRas-related protein Rab-22A1.32GTPRas family
1Z2ARas-related protein Rab-231.90GDPRas family
2OILRas-related protein Rab-252.30GDPRas family
1Z0ARas-related protein Rab-2A2.12GDPRas family
2A5JRas-related protein Rab-2B1.501GDPRas family
3DZ8Ras-related protein Rab-3B1.90GDPRas family
2GF9Ras-related protein Rab-3D1.53GDPRas family
2HUPRAS-related protein RAB-432.05GDPRas family
2BMDRAS-RELATED PROTEIN RAB4A1.80GDPRas family
2O52Ras-related protein Rab-4B2.20GDPRas family
1N6KRas-related protein Rab-5A1.55GDPRas family
2E9SRas-related protein Rab-6B1.78GDPRas family
1T91Ras-related protein Rab-71.90GTPRas family
4LHVRas-related protein Rab-8A1.95GDPRas family
1WMSRas-related protein Rab-9A1.25GDPRas family
4QXARas-related protein Rab-9A2.30GTPRas family
1U8ZRas-related protein Ral-A1.50GDPRas family
3X1WRas-related protein Rap-1b1.20GDPRas family
2FN4Ras-related protein R-Ras1.65GDPRas family
2ERYRas-related protein R-Ras21.70GDPRas family
4MITRho family GTPase2.35GTPRas family
3REFRho-like small GTPase1.95GDPRas family
2CLSRho-related GTP-binding protein RHO62.31GTPRas family
2FV8Rho-related GTP-binding protein RhoB1.90GDPRas family
2J1LRho-related GTP-binding protein RHOD2.50GDPRas family
1M7BRnd3/RhoE small GTP-binding protein2.00GTPRas family
2BCGSecretory pathway GDP dissociation inhibitor1.48GDPRas family
2EFHSimilarity to vacuolar protein sorting-associated protein VPS92.10GDPRas family
3BFKSmall GTPase Rab111.80GDPRas family
5C4MTransforming protein RhoA1.30GDPRas family
5C2KTransforming protein RhoA, Rac GTPase-activating protein 11.42GDPRas family
6EWZGTP pyrophosphokinase2.24GTPRegion found in RelA/SpoT proteins
3O0QRibonucleoside-diphosphate reductase1.80GDPRibonucleotide reductase, all-alpha domain
2CVWRibonucleoside-diphosphate reductase large chain 12.40GDPRibonucleotide reductase, barrel domain
5CA8Protein SEY12.30GDPRoot hair defective 3 GTP-binding protein (RHD3)
2RCNProbable GTPase EngC2.25GDPRsgA GTPase
2YV5YjeQ protein1.90GDPRsgA GTPase
4Z54Neuronal-specific septin-31.83GDPSeptin
4KV9Septin1.93GDPSeptin
5CYOSeptin-92.04GDPSeptin
2FH5Signal recognition particle receptor alpha subunit2.45GTPSignal recognition particle receptor beta subunit
1NRJSignal recognition particle receptor alpha subunit homolog1.70GTPSignal recognition particle receptor beta subunit
2IYLCell division protein FTSY2.10GDPSRP54-type protein, GTPase domain
5L3VSignal recognition particle 54 kDa protein2.30GDPSRP54-type protein, GTPase domain
2C03Signal recognition particle receptor1.24GDPSRP54-type protein, GTPase domain
3E70Signal recognition particle receptor1.97GDPSRP54-type protein, GTPase domain
5L3WSignal recognition particle receptor FtsY2.40GDPSRP54-type protein, GTPase domain
2ZGYPlasmid segregation protein parM1.90GDPStbA protein
4IENPutative acyl-CoA hydrolase2.00GDPThioesterase superfamily
1OFUCell division protein FtsZ2.10GDPTubulin/FtsZ family, GTPase domain
2R6RCell division protein ftsZ1.70GDPTubulin/FtsZ family, GTPase domain
2RHLCell Division Protein ftsZ2.45GDPTubulin/FtsZ family, GTPase domain
2VAPCell division protein FtsZ1.70GDPTubulin/FtsZ family, GTPase domain
4B46Cell division protein ftsZ1.90GDPTubulin/FtsZ family, GTPase domain
5XDTCell division protein FtsZ1.30GDPTubulin/FtsZ family, GTPase domain
4EI7Plasmid replication protein RepX1.90GDPTubulin/FtsZ family, GTPase domain
5IYZTubulin alpha-1B chain1.80GTPTubulin/FtsZ family, GTPase domain
2BTOTUBULIN BTUBA2.50GTPTubulin/FtsZ family, GTPase domain
3CB2Tubulin gamma-1 chain2.30GDPTubulin/FtsZ family, GTPase domain
3ZIDTUBULIN/FTSZ, GTPASE2.00GDPTubulin/FtsZ family, GTPase domain
4XCQTubZ2.39GDPTubulin/FtsZ family, GTPase domain
1RA7Genome polyprotein2.35GTPViral RNA-dependent RNA polymerase
1UVKRNA-dependent RNA polymerase2.45GTPViral RNA-dependent RNA polymerase
5XE0Genome polyprotein2.30GTPViral RNA-dependent RNA polymerase
3N6MRNA-dependent RNA polymerase2.50GTPViral RNA-dependent RNA polymerase
4UCIRNA-dependent RNA polymerase L2.21GTPVirus-capping methyltransferase
5KWKGalactoside 2-alpha-L-fucosyltransferase1.90GDPXyloglucan fucosyltransferase
2GJ8tRNA modification GTPase trmE1.70GDP50S ribosome-binding GTPase
2IRXDNA ligase-like protein Rv0938/MT09651.80GTPDNA primase small subunit
5KSPMitochondrial Rho GTPase 12.16GDPRas family
5KU1Mitochondrial Rho GTPase 12.50GDPRas family
5KUTMitochondrial Rho GTPase 21.69GDPRas family
3ZBQPHIKZ0391.70GDPTubulin/FtsZ family, GTPase domain
3R4VPutative uncharacterized protein1.67GDPTubulin/FtsZ family, GTPase domain
3DZHADP-ribosyl cyclase 11.60GTP-
4XJ3Cyclic AMP-GMP synthase1.65GTP-
3T34Dynamin-related protein 1AA2.41GDP-
1MREIGG2B-KAPPA JEL103 FAB (LIGHT CHAIN)2.30GDP-
4XULmg6622.26GTP-
5GOFMitofusin-11.60GTP-
5X6ZmRNA capping enzyme P52.10GDP-
4GMUPhosphoenolpyruvate carboxykinase1.20GTP-
3WNCProtein translation elongation factor 1A1.90GDP-
2QU8Putative nucleolar GTP-binding protein 12.01GDP-
5CK4Putative signal recognition particle protein1.89GDP-
4KU4Ras-3 from Cryphonectria parasitica1.60GDP-
3SFVRas-related protein Rab-1A1.73GDP-
2RHDSmall GTP-binding protein rab1a2.06GDP-
1JLRUracil Phosphoribosyltransferase2.45GTP-
a GTP, guanosine-5′-triphosphate; GDP, guanosine-5′-diphosphate; and GMP, guanosine-5′-monophosphate. b Family of proteins according to the Pfam classification [34]. Dashed line “-“ indicates that no knowledge of Pfam classification for the protein is available.
Table 2. Summary of the hydrogen bond patterns between guanine and the surrounding residues and their associated sequence motifs.
Table 2. Summary of the hydrogen bond patterns between guanine and the surrounding residues and their associated sequence motifs.
S. N.MotifHydrogen Bonding Pattern aPercentage/No. of Complexes
1.NKXDNi-Ki+1-Xi+2-Di+319.1%: 57 out of 298
2.NKXDNi-Ki+1-Xi+2-Di+319.1%: 57 out of 298
3.NKXDNi-Ki+1-Xi+2-Di+36.7%: 20 out of 298
4.NKXDNi-Ki+1-Xi+2-Di+311.7%: 35 out of 298
5.Non-NKXDDi/Ei plus20.5%: 61 out of 298
6.Non-NKXDOthers22.8%: 68 out of 298
a. The uppercase letters represent the NKXD motif, and “i” in the subscript represents the residue number. The amino acid that forms the hydrogen bond with the guanine ring is colored in red text with a bold face. The cut-off distance for hydrogen bonds is 3.5 Å.
Table 3. Pairwise interaction energies for representative intermolecular cation–π interacting pairs.
Table 3. Pairwise interaction energies for representative intermolecular cation–π interacting pairs.
S. NCation–π
Interaction Pair a
PDB
ID
Distance b
(Å)
Δ E I n t g a s
(Kcal/mol) c
Δ E D e h
(Kcal/mol)
Δ E I n t a q  d
(Kcal/mol)
1.G…K1311G7S3.44−10.740.13−10.61
2.G…K1242BMD4.25−6.07−4.23−10.30
3.G…R1301S4O3.42−8.51−1.30−9.81
4.G…K2053P324.26−6.49−2.89−9.38
5.G…R521RYA2.77−33.6524.57−9.08
6.G…K1261T913.87−4.05−4.76−8.81
7.G…K552IRX2.97−36.3828.30−8.09
8.G…R1425A073.57−6.38−1.56−7.94
9.G…R2176B9F2.87−28.5823.07−5.51
10.G…R1814B2P3.73−9.124.59−4.53
11.G…R902DYK4.29−2.780.72−2.06
12.G…K452V404.92−11.49.89−1.51
a “G” represents guanine. b The intermolecular distance between the positively charged residue and the guanine base. c Gas-phase interaction energies calculated at the B2PLYP-D3/cc-pVDZ level of theory. d Solution-phase interaction energies were calculated according to the equation Δ E I n t a q = Δ E I n t g a s + Δ E D e h , as described in the Theory and Methods section.
Table 4. Pairwise interaction energies for representative π–π stacking interaction pairs.
Table 4. Pairwise interaction energies for representative π–π stacking interaction pairs.
S. Nπ–π Stacking Pair aPDB IDAngle
(Degrees)
Distance b
(Å)
Δ E I n t g a s
(Kcal/mol) c
Δ E D e h
(Kcal/mol)
Δ E I n t a q
(Kcal/mol) d
1.G…F31RYA18.693.29−4.58−1.99−6.57
2.G…Y1613R4V10.943.53−5.36−1.03−6.39
3.G…W1893DZH2.953.40−7.131.08−6.05
4.G…Y6301UVK8.393.46−3.59−2.38−5.97
5.G…F243EVD8.703.30−4.38−0.99−5.37
6.G…Y945IGI20.393.51−9.696.14−3.55
7.G…W3594Q469.773.40−4.130.61−3.52
8.G…F1601JE153.833.54−3.06−0.18−3.24
9.G…F283KKQ77.413.98−2.40.84−1.56
10.G…F2275VYR56.854.04−2.751.73−1.02
11.G…F2936B9F78.433.96−1.070.26−0.81
12.G…F2771RPN15.533.46−0.670.04−0.63
13.G…Y3444XUL12.695.28−3.292.92−0.37
14.G…F1904LPS46.473.48−1.681.34−0.34
a “G” represents guanine. b The intermolecular distance between the aromatic residue and the guanine base. c Gas-phase interaction energies calculated at the B2PLYP-D3/cc-pVDZ level of theory. d Solution-phase interaction energies calculated according to the equation Δ E I n t a q = Δ E I n t g a s + Δ E D e h , as described in the Theory and Methods section.
Table 5. The interaction energies for various modes of non-bonded interactions between the guanine base and its surrounding residues in p21-ras (PDB ID: 1QRA).
Table 5. The interaction energies for various modes of non-bonded interactions between the guanine base and its surrounding residues in p21-ras (PDB ID: 1QRA).
S. N.Intermolecular
Pair a
Interaction Mode b Δ E I n t g a s
(Kcal/mol) c
Δ E D e h
(Kcal/mol)
Δ E I n t a q
(Kcal/mol) d
1.G…N116HB−2.471.98−0.49
2.G…K117mHB−6.203.62−2.58
3.G…D119HB−28.5924.51−4.08
4.G…A146mHB−8.936.74−2.19
5.G…K117Cation–π−5.66−5.15−10.81
6.G…K147Cation–π−4.43−2.96−7.39
7.G…F28π–π−4.431.44−2.99
a “G” represents guanine; the superscript “m” designates the main chain of the residue. b HB stands for hydrogen bond interaction. c The gas-phase interaction energies were calculated at the B2PLYP-D3/cc-pVDZ level of theory. d The solution-phase interaction energies were calculated according to the equation Δ E I n t a q = Δ E I n t g a s + Δ E D e h , as described in the Theory and Methods section.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bhatta, P.; Hu, X. Molecular Determinants for Guanine Binding in GTP-Binding Proteins: A Data Mining and Quantum Chemical Study. Int. J. Mol. Sci. 2024, 25, 12449. https://doi.org/10.3390/ijms252212449

AMA Style

Bhatta P, Hu X. Molecular Determinants for Guanine Binding in GTP-Binding Proteins: A Data Mining and Quantum Chemical Study. International Journal of Molecular Sciences. 2024; 25(22):12449. https://doi.org/10.3390/ijms252212449

Chicago/Turabian Style

Bhatta, Pawan, and Xiche Hu. 2024. "Molecular Determinants for Guanine Binding in GTP-Binding Proteins: A Data Mining and Quantum Chemical Study" International Journal of Molecular Sciences 25, no. 22: 12449. https://doi.org/10.3390/ijms252212449

APA Style

Bhatta, P., & Hu, X. (2024). Molecular Determinants for Guanine Binding in GTP-Binding Proteins: A Data Mining and Quantum Chemical Study. International Journal of Molecular Sciences, 25(22), 12449. https://doi.org/10.3390/ijms252212449

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop