Interactions between Endohedral Metallofullerenes and Proteins: The

1 hour ago - Endohedral metallofullerenes (EMFs) have great potential as radioisotope carriers for nuclear medicine and as contrast agents for X-ray a...
0 downloads 0 Views 4MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2018, 3, 13782−13789

http://pubs.acs.org/journal/acsodf

Interactions between Endohedral Metallofullerenes and Proteins: The Gd@C60−Lysozyme Model Fabio Bologna, Edoardo Jun Mattioli, Andrea Bottoni, Francesco Zerbetto, and Matteo Calvaresi*

Downloaded via 37.230.212.158 on October 27, 2018 at 20:37:27 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Dipartimento di Chimica “Giacomo Ciamician”, Alma Mater StudiorumUniversità di Bologna, via F. Selmi 2, 40126 Bologna, Italy ABSTRACT: Endohedral metallofullerenes (EMFs) have great potential as radioisotope carriers for nuclear medicine and as contrast agents for X-ray and magnetic resonance imaging. EMFs have still important restrictions for their use due to low solubility in physiological environments, low biocompatibility, nonspecific cellular uptake, and a strong dependence of their peculiar properties on physiological parameters, such as pH and salt content. Conjugation of the EMFs with proteins can overcome many of these limitations. Here we investigated the thermodynamics of binding of a model EMF (Gd@C60) with a protein (lysozyme) that is known to act as a host for the empty fullerene. As a rule, even if the shape of an EMF is exactly the same as that of the related fullerene, the interactions with a protein are significantly different. The estimated interaction energy (ΔGbinding) between Gd@C60 and lysozyme is −18.7 kcal mol−1, suggesting the possibility of using proteins as supramolecular carriers for EMFs. π−π stacking, hydrophobic interactions, surfactantlike interactions, and electrostatic interactions govern the formation of the hybrid between Gd@C60 and lysozyme. The comparison of the energy contributions to the binding between C60 or Gd@C60 and lysozyme suggests that, although shape complementarity remains the driving force of the binding, the presence of electron transfer from the gadolinium atom to the carbon cage induces a charge distribution on the fullerene cage that strongly affects its interaction with the protein.



INTRODUCTION Fullerenes are spherical molecules made only of carbon atoms. Atoms and molecular clusters can be encapsulated inside fullerenes to form endohedral fullerenes.1 The encapsulation of metal atoms creates a new class of hybrid molecules, which are called endohedral metallofullerenes (EMFs).1−5 The physicochemical properties of these hybrid materials differ from those of empty fullerenes because of the presence of electron transfer from the metal atoms to the carbon cage.1−5 The number of atoms encapsulated inside the cage defines the type of EMFs that can labeled mono-EMFs,1−5 di-EMFs,1−5 tri-EMFs,1−6 and cluster-EMFs.1−8 EMFs can find applications in a variety of fields that span molecular electronics, photovoltaics, materials science, and nanomedicine.1 A further interesting potential application is the exploitation of EMFs as contrast agents (CAs) for magnetic resonance imaging (MRI) and X-ray imaging as well as radioisotope carriers in nuclear medicine.9−14 Many metal ions that have found application in medicine15 are toxic and cannot be administered safely in vivo. Chelating agents are often used to coordinate the metal ions and reduce their toxicity.16 However, in vivo, these chelate complexes are prone to metal dissociation.17 The release of free metal is impossible when using EMFs. EMFs therefore provide a truly unique alternative to chelating compounds because of their resistance to metabolism and their high kinetic stability. Thus, incarceration of atoms inside fullerene cages prevents direct binding of the toxic metal ions with serum components and © 2018 American Chemical Society

tissues. A class of EMFs very promising for applications as MRI contrast agents (CAs) is represented by gadofullerenes.18−27 Gadofullerenes are paramagnetic compounds that entrap Gd3+ ions. They offer fundamental advantages over conventional chelate-based MRI CAs. The aquated Gd3+ ion is toxic,17 and its toxicity is usually sequestered by chelation with multidentate ligands.16 Both linear and macrocyclic chelating ligands were exploited, and several compounds are now available.16 However, even the best-designed chelates are characterized by some thermodynamic instability and the associated metabolic-ion release. A possible risk associated with the administration of these contrast agent is the acute nephrotoxicity.17 In particular, nephropatic patients could develop nephrogenic systemic fibrosis after exposure to gadolinium, a potentially fatal disease.17 The encapsulation of metals inside the fullerene cages cannot be considered as a chelation, i.e., a coordinative, dative bonding. The Gd3+ metal ion is trapped inside the biologically stable fullerene cage,1 so that the toxic Gd3+ ion is never released in vivo. This obvious advantage of gadofullerene structures is especially important for CAs that require longer retention times in vivo, such as those for blood-pool imaging. Gadofullerene species have recently been shown to exhibit T1 relaxivity up to 500 times greater27 than those of CAs presently Received: August 3, 2018 Accepted: October 9, 2018 Published: October 22, 2018 13782

DOI: 10.1021/acsomega.8b01888 ACS Omega 2018, 3, 13782−13789

ACS Omega

Article

charges were calculated for the most stable structure of Gd@C60 according to Merz−Kollman (MK) scheme (Figure 1).

in clinical use. These new nanoscale materials are the highest performing MRI CA known to date. The carbon cage itself plays an important role in the large value of the relaxivity of the gadofullerenes. In the relaxation process, the gadofullerene can simultaneously relax the protons of many hydrogen-bonded water molecules on their ∼200 Å2 paramagnetic surface.21 The increased contrast and stable encapsulation of Gd3+ potentially offer the possibility of using lower dosages of more potent and safer paramagnetic contrast agents, as compared with the performance of clinically used Gd chelates. Although gadofullerenes overcome much of the limitations of current CAs, demonstrating superior relaxivity properties and no Gd3+ release, they still present important restrictions in their use: (i) low biocompatibility, (ii) difficulty to direct or target the gadofullerenes to specific cells or tissues, (iii) proton relaxivities that display a remarkable pH dependency,24 (iv) fullerene-based materials show aggregation phenomena that depend on the salt content of the medium and, therefore, affect their behavior in biological or medical applications,25 (v) short retention times. Conjugation of EMFs with proteins can overcome most, or all, of these limitations. The ability of fullerenes to interact with proteins was demonstrated in many systems, both experimentally28−43 and computationally.43−53 A few noteworthy advantages of protein EMF bioconjugates are (a) a EMF-bioconjugate can be perfectly biocompatible;54 (b) the relaxivity will not depend on pH or salt concentration, since the EMF-bioconjugate does not aggregate;28 (c) specific tissue targeting becomes possible with binding to specific proteins or antibodies.55−57 Crucially, conjugation of a paramagnetic gadofullerene and a protein can improve relaxivity, since the interaction with proteins can slow down the molecular rotation of the EMF and result in a relaxivity increase.58 In the past, molecular dynamics (MD) simulations have been carried out to investigate the interactions between gadofullerenol and proteins.59−61 The aim of this work is to investigate the possibility of conjugating pristine EMFs with proteins in a noncovalent way. Gd@C60− lysozyme was chosen as a representative case because it is possible to make a comparison with the C60−lysozyme system, which has been well investigated both experimentally and computationally.28,30,51,52,54 The analysis of the energy contributions of the binding between Gd@C60 and lysozyme also supplies guidelines for understanding protein−EMF interactions.

Figure 1. (a) Optimized structure of Gd@C60. (b) Charge distribution on the Gd@C60 cage; in blue negative charges, in red positive charges.

Docking Gd@C60 to Lysozyme. The docking procedure provides 27 privileged positions of Gd@C60 interacting with lysozyme (Figure 2a). To rank the different poses, the molecular mechanics-generalized born surface area (MMGBSA) value of the interaction63 between Gd@C60 and lysozyme, in the minimized structures, was used as a scoring



RESULTS AND DISCUSSION Density Functional Theory (DFT) Calculations on Gd@C60 Molecules. Due to the lack of crystallographic structures for Gd@C60, to find the most stable state and geometry of Gd@C60, the Gd ion was systematically inserted into the crystallographic structure of C60 in different positions inside the cage, and a series of geometry optimization calculations were carried out, using different multiplicities. In agreement with previous calculations,62 it was found that the Gd ion is placed close to one of the hexagonal faces of the C60 cage. The total spin multiplicity of Gd@C60 is S = 7. ESP

Figure 2. (a) Interaction sites of Gd@C60 with lysozyme. (b) In red, Gd@C60 in the most favorite binding pocket, in gray Gd@C60 in the C60 binding pocket. 13783

DOI: 10.1021/acsomega.8b01888 ACS Omega 2018, 3, 13782−13789

ACS Omega

Article

Table 1. Largest Contributions (ΔGbinding > 2.0 kcal mol−1) to ΔGbinding of Individual Residues (kcal mol−1) in the Most Favorite Binding Sitea

function. The position that maximizes the interaction of Gd@C60 (Figure 2b) provides the initial coordinates for a 200 ns MD simulation of the adduct in an explicit water box. To estimate the binding energy between lysozyme and Gd@C60 and their components, we carried out a MM-GBSA analysis of the trajectory. This methodology has provided valuable information to estimate the interaction between proteins and carbon nanomaterials.64−68 Gd@C 60 Binding Pocket. The interaction energy (ΔGbinding) between Gd@C60 and lysozyme, in the Gd@C60 most favorite binding pocket, was evaluated to be −18.7 kcal mol−1, a value very close to the C60−lysozyme interaction (−18.5 kcal mol−1).51 Experimentally, lysozyme was used as a supramolecular host for C60.28,30,54 These results indicate the feasibility of the noncovalent bioconjugation of EMFs with proteins. Analysis of the binding components of the energy (Figure 3) shows that, even if the Gd@C60 is highly charged, van der Waals interactions remain the driving force for the binding (−41.1 kcal mol−1).

Cys 30 Lys 33 Phe 34 Arg 114 Cys 115 a

−2.7 −2.7 −2.9 −6.6 −2.4

Thr 118 Asp 119 Val 120 Ala 122 Trp 123

−3.1 −2.4 −4.0 −2.6 −6.9

In bold values greater than 4 kcal mol−1.

Figure 3. Energy components of ΔGbinding of Gd@C60 with lysozyme in the most favorite binding site.

Figure 4. (a) π−π stacking interactions, (b) hydrophobic interactions, (c) surfactant-like interactions, and (d) electrostatic interactions between lysozyme residues and Gd@C60.

Electrostatic interactions account for only −3.0 kcal mol−1. In the Gd@C60, the electron transfer due to the presence of the Gd ion inside the cage induces, at the same time, positive and negative charges on the fullerene cage, generating a plethora of stabilizing and destabilizing terms that counterbalance each other. The fact that Gd@C60 maintains its hydrophobic character is proved by the negative value of the hydrophobic interactions, i.e., nonpolar solvation, that assist the binding (−3.7 kcal mol−1). This value is smaller than that of the vdW interactions. Polar solvation (12.6 kcal mol−1) and entropy contribution (16.5 kcal mol−1) are detrimental to the binding, since their contributions are positive. The per-residue decomposition of ΔGbinding helps in understanding the chemical origin of the various contributions. Proteins are able to establish a large number of interactions with carbon nanoparticles that include π−π stacking, hydrophobic interactions, surfactant-like interactions, and electrostatic interactions.69−73 Table 1 shows that a variety of interactions are present in the protein−EMF adduct. π−π stacking between aromatic residues and the conjugated surface of carbon nanomaterials usually governs their interaction.69−74 For Gd@C60, π-stacking contacts between the cage and the indolic group of Trp 123 (sandwich-like) and the phenyl group of Phe 34 (T-shaped) are important stabilizing terms (Figure 4a). Hydrophobic interactions are established between aliphatic residues and the Gd@C60 surface. Amino acids with a

hydrophobic side chain tend to bind to the surface of Gd@C60, reducing the interfacial energy with the water interface (hydrophobic effect).69 Usually, the interaction energy increases with an increase in the exposed hydrophobic surface of the protein residue.69 In the present case, the longer aliphatic chain of Val 120 interacts more strongly with the Gd@C60 than Ala 122 (Figure 4b). The interactions of Cys 30 and Cys 115 with Gd@C60 deserve special attention. These residues form a disulfide bond that interacts directly with Gd@C60 (Figure 4b). The importance of this kind of interaction was recently highlighted by Hirano and co-workers for lysozyme and carbon nanotubes.75,76 Our computations evidenced surfactant-like interactions between the Thr 118 and Lys 33 residues and Gd@C60 (Figure 4c).69,77,78 The hydrophobic aliphatic chains of these residues interact with Gd@C60, whereas the hydrophilic groups head toward water. In the interaction between carbon nanoparticles and proteins, electrostatic interactions are generally quantitatively smaller than those previously discussed. Positively charged residues (arginine and lysine, in particular) can interact with the carbon nanoparticle surfaces via cation−π interactions. When a net charge is not present on the residue, a chargetransfer mechanism may still occur, for example, during physisorption of aminic and amidic residues. In contrast, when we consider EMFs, where net charges exist on the fullerene cage, these interactions can have significant values. 13784

DOI: 10.1021/acsomega.8b01888 ACS Omega 2018, 3, 13782−13789

ACS Omega

Article

Figure 5. (a) Binding of Gd@C60 in the C60 binding pocket. (b) Binding of C60 in the C60 binding pocket Gd@C60. (c) Total binding energy (ΔGbinding) and energy components of ΔGbinding of Gd@C60 (in gray) and C60 (in black) with lysozyme, in the C60 binding pocket.

distribution on the fullerene cage (Figure 1), generating an additional stabilizing term in the binding energy with lysozyme (Gd@C60 electrostatic term = −1.8 kcal mol−1). The appearance of this term is extremely important because it explains qualitatively many of the differences observed with Gd@C60 and deserves an in-depth analysis. Even if the global effect is low, it is the result of many contributions that counterbalance each other. In a dynamical picture, one can see the effect of these stabilizing and destabilizing terms as a sequence of continuous kicks toward the Gd@C60 cage. This increases the mobility of the Gd@C60, when compared to that of the C60, especially in a crevice-like binding pocket, such as the substrate binding pocket of lysozyme (Figure 6).

Analysis of the most interacting residues shows that Arg 114, Lys 33, and Asp 199 strongly interact with Gd@C60. Interestingly, in the Gd@C60 cage, there is a dipolar distribution of charges (Figure 1b). In fact, the Gd ion is positioned close to one of the hexagonal faces of the C60 cage. This induces negative charges on the closest atoms via electron transfer and polarizes the Gd@C60 cage. In the hemisphere opposite to the Gd binding position, positive charges are induced. On average, the fullerene cage in Gd@C60 is approximately half positive and half negative. This charge distribution is mirrored in the Gd@C60 binding pocket, where positively charged Arg 114 and Lys 33 face the negative hemisphere of the fullerene, and on the opposite side, Asp 199 faces the positive charges (Figure 4d), generating strongly stabilizing Coulombic interactions between the protein binding pocket and Gd@C60. Gd@C60 vs C60 in the C60 Binding Pocket. In the previous section, we observed that the presence of a Gd atom encapsulated in the C60 cage changes the favorite recognition pocket for the binding to lysozyme. To compare directly the differences in the protein binding between C60 and Gd@C60, we estimated the total binding energy and their contributions, when the two fullerenes are bound in the same protein pocket (Figure 5). The C60 binding pocket is well characterized, and the corresponding binding thermodynamics has already been investigated both experimentally and computationally;28,30,51,52,54 the same binding pocket was used for Gd@C60. Thus, MD simulations were carried out for the Gd@C60− lysozyme adduct, when it is bound in the C60 binding pocket (Figure 5a). In this case, the interaction energy (ΔGbinding) between Gd@C60 and lysozyme is −4.8 kcal mol−1, a lower value when compared to the interaction energy of C60 with lysozyme (−18.5 kcal mol−1). This result confirms the predictive accuracy of the docking protocol developed for EMFs. The analysis of the binding components of the energy (Figure 5c) of Gd@C60 in the C60 binding site provides an explanation of the different behavior of C60 and Gd@C60 in their interaction with proteins. The first difference is the presence of a new term for Gd@C60, corresponding to the electrostatic interaction. Whereas the highly symmetric C60 molecule lacks net charges on the carbon atoms, for the Gd@C60 cage, the presence of the electron transfer from the gadolinium atom to the carbon cage induces a charge

Figure 6. Trajectories in the C60 binding site during the 200 ns MD simulation of the (a) Gd@C60 cage (Gd ion is removed for clarity) and (b) the C60 cage. 13785

DOI: 10.1021/acsomega.8b01888 ACS Omega 2018, 3, 13782−13789

ACS Omega

Article

it because of the net charges present in the Gd@C60 cage. Only if there is an ideal pocket able to accommodate the Gd@C60 cage can the binding with protein take place.

In the case of C60 (Figure 6b), protein residues stick to its surface to maximize vdW interactions (C60 vdW contribution = −45.1 kcal mol−1) and become glued to the fullerene cage (C60 entropic term = 18.1 kcal mol−1). In contrast, the binding of the Gd@C60 to the C60 binding pocket is characterized by higher mobility (Figure 6a), due to the charges that are present in the Gd@C60 cage. The enhanced mobility of both Gd@C60 and protein residues interacting with the fullerene cage reduces the entropic penalty of Gd@C60 upon lysozyme binding (Gd@C60 entropic term = 13.8 kcal mol−1), but at the same time decreases the van der Waals interactions, which are the driving force for the interaction between Gd@C60 and lysozyme (Gd@C60 vdW contribution = −26.2 kcal mol−1). Hydrophobic interactions, i.e., nonpolar solvation, assist the binding, even if the corresponding value for Gd@C 60 (Gd@C60 nonpolar solvation term = −2.8 kcal mol−1) is smaller than that for the empty hydrophobic C60 cage (C60 nonpolar solvation term = −4.3 kcal mol−1). The polar solvation term is detrimental for the binding for both C60 and Gd@C60, but the penalty is smaller for Gd@C60 than C60. The binding of the fullerene cage occurs in the substrate binding pocket of lysozyme, that is a region exposed to water and where amino acids with polar side groups are located. The hydrophilic parts of these residues, upon formation of the complex with fullerenes, are forcefully desolvated, causing a destabilization of the system. For the hydrophobic C60 cage (C 60 polar solvation term = 12.8 kcal mol −1 ), this destabilization is greater when compared to that of the charged cage of Gd@C60 (Gd@C60 polar solvation term = 12.1 kcal mol−1). To understand the differences between C60 and Gd@C60, upon binding, it is interesting to re-analyze the solvation binding terms, from the point of view of C60 and Gd@C60 (Figure 7).



CONCLUSIONS The calculated interaction energy between Gd@C60 and lysozyme is evaluated to be −18.7 kcal mol−1, suggesting the possibility of using proteins as supramolecular carriers for EMFs. π−π stacking (with Trp 123, Phe 34), hydrophobic interactions (with Val 120, Ala 122, Cys 30, Cys 115), surfactant-like interactions (with Thr 118, Lys 33), and electrostatic interactions (with Arg 114, Lys 33, Asp 199) govern the global interaction between lysozyme and Gd@C60. vdW interactions and shape complementarity remain the driving forces for the binding, even if the presence of the electron transfer from the gadolinium atom to the carbon cage induces a charge distribution on the fullerene cage that strongly affects its interaction with lysozyme. In fact, a different protein binding pocket, is identified for Gd@C60 with respect to C60. Although C60 has a natural tendency to bind to proteins, Gd@C60 prefers to remain in water because, upon binding, a positive polar solvation term (which is null for C60) exists and represents an energy penalty. In a dynamical picture, the presence of many stabilizing and destabilizing electrostatic terms that are located closely to each other causes continuous kicks to the Gd@C60 cage, increasing its mobility. In conclusion, only the interaction with a well-suited binding pocket, showing charge complementarity to the EMF and a well-defined complementary shape, allows EMF binding to the protein. The presence of net charges on the cage of EMF increases the constraints to make a protein binding pocket suitable for EMF and makes its identification more complicated.



COMPUTATIONAL DETAILS QM Calculation on Gd@C60 Molecules. The electronic structure and geometry of the Gd@C60 adduct were investigated by means of DFT calculations using the Gaussian 09 series of programs.79 Full geometric optimizations for Gd@C60 were carried out. A benchmark to evaluate the performance of different DFT functionals and basis sets to reproduce experimental data on gadofullerenes was recently performed by Zhou and co-workers.80 Following their results, the pure generalized gradient approximation functional Perdew−Burke−Ernzerhof81 was used for DFT calculations. To take into account the scalar relativistic effects, the effective core potential triple split basis set (CEP-121G)82 was used to describe the Gd atom, whereas for the carbon atoms, we employed the 6-31G* basis set.79 When pure functionals are used, the common medium-sized double-split 6-31G* is good enough to calculate reasonable septet−nonet gap and Gd−C distance.80 Frequency calculations were also performed after geometry optimization to ensure that the obtained structures are real minima on the potential energy surface. Parameterization of the Gd@C 60 Molecule for Molecular Mechanics. The fullerene cage was modeled using the CA atom type (sp2 Aromatic Carbon parameter). ESP charges were calculated for the most stable structure of Gd@C60 according to the Merz−Kollman (MK) scheme, a protocol used previously to study Gd−DOTA complexes.83 van der Waals parameters for the Gd atom were taken from the same study.83

Figure 7. Total solvation energy and polar and nonpolar energy components of total solvation energy for Gd@C60 (in gray) and C60 (in black) with lysozyme, in the C60 binding site.

It is clear that the difference is due to the presence of charges in the Gd@C60 cage. The binding between C60 and a protein is always favored because C60 is a purely hydrophobic sphere. When it binds to a protein binding pocket, there is always a stabilizing nonpolar term, whereas the polar solvation is null (net charges on the cage are zero). The nonpolar solvation term takes also into account the entropy increase due to the water molecules in the first hydration shell that are tightly bound to the protein and are set free upon noncovalent interactions with the fullerene (hydrophobic effect). On the opposite, for Gd@C60, even if a stabilizing nonpolar term is still present, the polar solvation term tends to overcome 13786

DOI: 10.1021/acsomega.8b01888 ACS Omega 2018, 3, 13782−13789

ACS Omega

Article

Docking Gd@C60. A docking protocol,44,77,84 recently validated for the study of interactions between proteins and nano-objects, was used to generate the coordinates of the adduct between lysozyme and Gd@C60. This methodology is based on a two-step protocol of docking. The shape complementarity algorithm of PatchDock85 was used to generate initial docking models of lysozyme and Gd@C60. The docking poses were then refined with FireDock.86 Sidechain flexibility is modeled by rotamers and Monte Carlo minimization. Following the rearrangement of the side chains, the relative position of the docking partners is refined by Monte Carlo minimization.87 All poses were fully minimized by AMBER 12.88 The ff12SB force field was used to model the proteins.88 The minimization was carried out with the Sander program, using the Generalized Born model for the solvation89 and no cut-off for van der Waals and electrostatic interactions was adopted. MD Production and MM-GBSA Analysis. The same protocol of equilibration, MD production, and molecular mechanics-generalized born surface area (MM-GBSA) recently used for C60−lysozyme complexes51 was used in this work. The overall sampling time was 200 ns. The calculation and decomposition of binding free energy, ΔGbind, between fullerenes and lysozyme were performed using the MMGBSA scheme.63 To obtain an estimate of the binding entropy, the translational and rotational entropies (standard statistical mechanical formulas), and the vibrational entropy contribution (normal-mode calculations, nmode program in AMBER) for the complex, receptor, and ligand are calculated during the MD trajectory. The results are averaged using the PTRAJ program via MMPBSA.py.



(6) Xu, W.; Feng, L.; Calvaresi, M.; Liu, J.; Liu, Y.; Niu, B.; Shi, Z.; Lian, Y.; Zerbetto, F. An Experimentally Observed Trimetallofullerene Sm3@Ih -C80: Encapsulation of Three Metal Atoms in a Cage without a Nonmetallic Mediator. J. Am. Chem. Soc. 2013, 135, 4187−4190. (7) Dunsch, L.; Yang, S. Endohedral Clusterfullerenesplaying with Cluster and Cage Sizes. Phys. Chem. Chem. Phys. 2007, 9, 3067−3081. (8) Yang, S.; Liu, F.; Chen, C.; Jiao, M.; Wei, T. Fullerenes Encaging Metal Clustersclusterfullerenes. Chem. Commun. 2011, 47, 11822− 11839. (9) Cagle, D. W.; Kennel, S. J.; Mirzadeh, S.; Alford, J. M.; Wilson, L. J. In Vivo Studies of Fullerene-Based Materials Using Endohedral Metallofullerene Radiotracers. Proc. Natl. Acad. Sci. U.S.A. 1999, 96, 5182−5187. (10) Iezzi, E. B.; Duchamp, J. C.; Fletcher, K. R.; Glass, T. E.; Dorn, H. C. Lutetium-Based Trimetallic Nitride Endohedral Metallofullerenes: New Contrast Agents. Nano Lett. 2002, 2, 1187−1190. (11) Shultz, M. D.; Duchamp, J. C.; Wilson, J. D.; Shu, C.-Y.; Ge, J.; Zhang, J.; Gibson, H. W.; Fillmore, H. L.; Hirsch, J. I.; Dorn, H. C.; Fatouros, P. P. Encapsulation of a Radiolabeled Cluster Inside a Fullerene Cage, 177LuxLu(3−x ) N@C80: An Interleukin-13-Conjugated Radiolabeled Metallofullerene Platform. J. Am. Chem. Soc. 2010, 132, 4980−4981. (12) Meng, J.; Wang, D.-L.; Wang, P. C.; Jia, L.; Chen, C.; Liang, X.J. Biomedical Activities of Endohedral Metallofullerene Optimized for Nanopharmaceutics. J. Nanosci. Nanotechnol. 2010, 10, 8610−8616. (13) Shultz, M. D.; Wilson, J. D.; Fuller, C. E.; Zhang, J.; Dorn, H. C.; Fatouros, P. P. Metallofullerene-Based Nanoplatform for Brain Tumor Brachytherapy and Longitudinal Imaging in a Murine Orthotopic Xenograft Model. Radiology 2011, 261, 136−143. (14) Chen, Z.; Ma, L.; Liu, Y.; Chen, C. Applications of Functionalized Fullerenes in Tumor Theranostics. Theranostics 2012, 2, 238−250. (15) Thompson, K. H.; Orvig, C. Boon and Bane of Metal Ions in Medicine. Science 2003, 300, 936−939. (16) Caravan, P.; Ellison, J. J.; McMurry, T. J.; Lauffer, R. B. Gadolinium(III) Chelates as MRI Contrast Agents: Structure, Dynamics, and Applications. Chem. Rev. 1999, 99, 2293−2352. (17) Rogosnitzky, M.; Branch, S. Gadolinium-Based Contrast Agent Toxicity: A Review of Known and Proposed Mechanisms. Biometals 2016, 29, 365−376. (18) Bolskar, R. D. Gadofullerene MRI Contrast Agents. Nanomedicine 2008, 3, 201−213. (19) Sitharaman, B.; Wilson, L. J. Gadofullerenes and Gadonanotubes: A New Paradigm for High-Performance Magnetic Resonance Imaging Contrast Agent Probes. J. Biomed. Nanotechnol. 2007, 3, 342−352. (20) Zheng, J.-P.; Zhen, M.-M.; Wang, C.-R.; Shu, C.-Y. Recent Progress of Molecular Imaging Probes Based on Gadofullerenes. Chin. J. Anal. Chem. 2012, 40, 1607−1615. (21) Mikawa, M.; Kato, H.; Okumura, M.; Narazaki, M.; Kanazawa, Y.; Miwa, N.; Shinohara, H. Paramagnetic Water-Soluble Metallofullerenes Having the Highest Relaxivity for MRI Contrast Agents. Bioconjug. Chem. 2001, 12, 510−514. (22) Tóth, É .; Bolskar, R. D.; Borel, A.; González, G.; Helm, L.; Merbach, A. E.; Sitharaman, B.; Wilson, L. J. Water-Soluble Gadofullerenes: Toward High-Relaxivity, pH-Responsive MRI Contrast Agents. J. Am. Chem. Soc. 2005, 127, 799−805. (23) Laus, S.; Sitharaman, B.; Tóth, É .; Bolskar, R. D.; Helm, L.; Wilson, L. J.; Merbach, A. E. Understanding Paramagnetic Relaxation Phenomena for Water-Soluble Gadofullerenes. J. Phys. Chem. C 2007, 111, 5633−5639. (24) Shu, C.-Y.; Zhang, E.-Y.; Xiang, J.-F.; Zhu, C.-F.; Wang, C.-R.; Pei, X.-L.; Han, H.-B. Aggregation Studies of the Water-Soluble Gadofullerene Magnetic Resonance Imaging Contrast Agent: [Gd@ C82O6(OH)16(NHCH2CH2COOH)8]X. J. Phys. Chem. B 2006, 110, 15597−15601. (25) Laus, S.; Sitharaman, B.; Tóth, É .; Bolskar, R. D.; Helm, L.; Asokan, S.; Wong, M. S.; Wilson, L. J.; Merbach, A. E. Destroying Gadofullerene Aggregates by Salt Addition in Aqueous Solution of

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Fabio Bologna: 0000-0002-5957-4776 Andrea Bottoni: 0000-0003-2966-4065 Francesco Zerbetto: 0000-0002-2419-057X Matteo Calvaresi: 0000-0002-9583-2146 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This study was supported by the Italian Ministry of Education, University and Research MIUR-SIR Programme no. RBSI149ZN9-BIOTAXI funded to M.C.



REFERENCES

(1) Popov, A. A.; Yang, S.; Dunsch, L. Endohedral Fullerenes. Chem. Rev. 2013, 113, 5989−6113. (2) Chaur, M. N.; Melin, F.; Ortiz, A. L.; Echegoyen, L. Chemical, Electrochemical, and Structural Properties of Endohedral Metallofullerenes. Angew. Chem., Int. Ed. 2009, 48, 7514−7538. (3) Rodríguez-Fortea, A.; Balch, A. L.; Poblet, J. M. Endohedral Metallofullerenes: A Unique Host−guest Association. Chem. Soc. Rev. 2011, 40, 3551−3563. (4) Yamada, M.; Akasaka, T.; Nagase, S. Endohedral Metal Atoms in Pristine and Functionalized Fullerene Cages. Acc. Chem. Res. 2010, 43, 92−102. (5) Lu, X.; Feng, L.; Akasaka, T.; Nagase, S. Current Status and Future Developments of Endohedral Metallofullerenes. Chem. Soc. Rev. 2012, 41, 7723−7760. 13787

DOI: 10.1021/acsomega.8b01888 ACS Omega 2018, 3, 13782−13789

ACS Omega

Article

Gd@C60(OH)x and Gd@C60[C(COOH2)]10. J. Am. Chem. Soc. 2005, 127, 9368−9369. (26) Shu, C.; Corwin, F. D.; Zhang, J.; Chen, Z.; Reid, J. E.; Sun, M.; Xu, W.; Sim, J. H.; Wang, C.; Fatouros, P. P.; Esker, A. R.; Gibson, H. W.; Dorn, H. C. Facile Preparation of a New Gadofullerene-Based Magnetic Resonance Imaging Contrast Agent with High 1H Relaxivity. Bioconjug. Chem. 2009, 20, 1186−1193. (27) Zhang, J.; Fatouros, P. P.; Shu, C.; Reid, J.; Owens, L. S.; Cai, T.; Gibson, H. W.; Long, G. L.; Corwin, F. D.; Chen, Z.-J.; Dorn, H. C. High Relaxivity Trimetallic Nitride (Gd3N) Metallofullerene MRI Contrast Agents with Optimized Functionality. Bioconjug. Chem. 2010, 21, 610−615. (28) Di Giosia, M.; Bomans, P. H. H.; Bottoni, A.; Cantelli, A.; Falini, G.; Franchi, P.; Guarracino, G.; Friedrich, H.; Lucarini, M.; Paolucci, F.; Rapino, S.; Sommerdijk, N. A. J. M.; Soldà, A.; Valle, F.; Zerbetto, F.; Calvaresi, M. Proteins as Supramolecular Hosts for C60: A True Solution of C60 in Water. Nanoscale 2018, 10, 9908−9916. (29) Di Giosia, M.; Valle, F.; Cantelli, A.; Bottoni, A.; Zerbetto, F.; Calvaresi, M. C60 Bioconjugation with Proteins: Towards a Palette of Carriers for All pH Ranges. Materials 2018, 11, 691. (30) Calvaresi, M.; Arnesano, F.; Bonacchi, S.; Bottoni, A.; Calò, V.; Conte, S.; Falini, G.; Fermani, S.; Losacco, M.; Montalti, M.; Natile, G.; Prodi, L.; Sparla, F.; Zerbetto, F. C60@Lysozyme: Direct Observation by Nuclear Magnetic Resonance of a 1:1 Fullerene Protein Adduct. ACS Nano 2014, 8, 1871−1877. (31) Friedman, S. H.; DeCamp, D. L.; Sijbesma, R. P.; Srdanov, G.; Wudl, F.; Kenyon, G. L. Inhibition of the HIV-1 Protease by Fullerene Derivatives: Model Building Studies and Experimental Verification. J. Am. Chem. Soc. 1993, 115, 6506−6509. (32) Miao, Y.; Xu, J.; Shen, Y.; Chen, L.; Bian, Y.; Hu, Y.; Zhou, W.; Zheng, F.; Man, N.; Shen, Y.; Zhang, Y.; Wang, M.; Wen, L. Nanoparticle as Signaling Protein Mimic: Robust Structural and Functional Modulation of CaMKII upon Specific Binding to Fullerene C60 Nanocrystals. ACS Nano 2014, 8, 6131−6144. (33) Mentovich, E.; Belgorodsky, B.; Gozin, M.; Richter, S.; Cohen, H. Doped Biomolecules in Miniaturized Electric Junctions. J. Am. Chem. Soc. 2012, 134, 8468−8473. (34) Zhen, M.; Zheng, J.; Ye, L.; Li, S.; Jin, C.; Li, K.; Qiu, D.; Han, H.; Shu, C.; Yang, Y.; Wang, C. Maximizing the Relaxivity of GdComplex by Synergistic Effect of HSA and Carboxylfullerene. ACS Appl. Mater. Interfaces 2012, 4, 3724−3729. (35) Ratnikova, T. A.; Nedumpully Govindan, P.; Salonen, E.; Ke, P. C. In Vitro Polymerization of Microtubules with a Fullerene Derivative. ACS Nano 2011, 5, 6306−6314. (36) Innocenti, A.; Durdagi, S.; Doostdar, N.; Amanda Strom, T.; Barron, A. R.; Supuran, C. T. Nanoscale Enzyme Inhibitors: Fullerenes Inhibit Carbonic Anhydrase by Occluding the Active Site Entrance. Bioorg. Med. Chem. 2010, 18, 2822−2828. (37) Durdagi, S.; Supuran, C. T.; Strom, T. A.; Doostdar, N.; Kumar, M. K.; Barron, A. R.; Mavromoustakos, T.; Papadopoulos, M. G. In Silico Drug Screening Approach for the Design of Magic Bullets: A Successful Example with Anti-HIV Fullerene Derivatized Amino Acids. J. Chem. Inf. Model. 2009, 49, 1139−1143. (38) Yang, S.-T.; Wang, H.; Guo, L.; Gao, Y.; Liu, Y.; Cao, A. Interaction of Fullerenol with Lysozyme Investigated by Experimental and Computational Approaches. Nanotechnology 2008, 19, No. 395101. (39) Zhang, X.; Shu, C.; Xie, L.; Wang, C.; Zhang, Y.; Xiang, J.; Li, L.; Tang, Y. Protein Conformation Changes Induced by a Novel Organophosphate-Containing Water-Soluble Derivative of a C60 Fullerene Nanoparticle. J. Phys. Chem. C 2007, 111, 14327−14333. (40) Pastorin, G.; Marchesan, S.; Hoebeke, J.; Da Ros, T.; EhretSabatier, L.; Briand, J.-P.; Prato, M.; Bianco, A. Design and Activity of Cationic Fullerene Derivatives as Inhibitors of Acetylcholinesterase. Org. Biomol. Chem. 2006, 4, 2556−2562. (41) Belgorodsky, B.; Fadeev, L.; Kolsenik, J.; Gozin, M. Formation of a Soluble Stable Complex between Pristine C60-Fullerene and a Native Blood Protein. ChemBioChem 2006, 7, 1783−1789.

(42) Belgorodsky, B.; Fadeev, L.; Ittah, V.; Benyamini, H.; Zelner, S.; Huppert, D.; Kotlyar, A. B.; Gozin, M. Formation and Characterization of Stable Human Serum Albumin−Tris-Malonic Acid [C60]Fullerene Complex. Bioconjug. Chem. 2005, 16, 1058−1062. (43) Liu, Y.; Yan, B.; Winkler, D. A.; Fu, J.; Zhang, A. Competitive Inhibition Mechanism of Acetylcholinesterase without Catalytic Active Site Interaction: Study on Functionalized C60 Nanoparticles via in Vitro and in Silico Assays. ACS Appl. Mater. Interfaces 2017, 9, 18626−18638. (44) Calvaresi, M.; Zerbetto, F. Baiting Proteins with C60. ACS Nano 2010, 4, 2283−2299. (45) Calvaresi, M.; Furini, S.; Domene, C.; Bottoni, A.; Zerbetto, F. Blocking the Passage: C60 Geometrically Clogs K+ Channels. ACS Nano 2015, 9, 4827−4834. (46) Turabekova, M.; Rasulev, B.; Theodore, M.; Jackman, J.; Leszczynska, D.; Leszczynski, J. Immunotoxicity of Nanoparticles: A Computational Study Suggests That CNTs and C60 Fullerenes Might Be Recognized as Pathogens by Toll-like Receptors. Nanoscale 2014, 6, 3488−3495. (47) Nedumpully Govindan, P.; Monticelli, L.; Salonen, E. Mechanism of Taq DNA Polymerase Inhibition by Fullerene Derivatives: Insight from Computer Simulations. J. Phys. Chem. B 2012, 116, 10676−10683. (48) Calvaresi, M.; Zerbetto, F. Fullerene Sorting Proteins. Nanoscale 2011, 3, 2873−2881. (49) Wu, X.; Yang, S.-T.; Wang, H.; Wang, L.; Hu, W.; Cao, A.; Liu, Y. Influences of the Size and Hydroxyl Number of Fullerenes/ Fullerenols on Their Interactions with Proteins. J. Nanosci. Nanotechnol. 2010, 10, 6298−6304. (50) Benyamini, H.; Shulman-Peleg, A.; Wolfson, H. J.; Belgorodsky, B.; Fadeev, L.; Gozin, M. Interaction of C60-Fullerene and Carboxyfullerene with Proteins: Docking and Binding Site Alignment. Bioconjug. Chem. 2006, 17, 378−386. (51) Calvaresi, M.; Bottoni, A.; Zerbetto, F. Thermodynamics of Binding Between Proteins and Carbon Nanoparticles: The Case of C60@Lysozyme. J. Phys. Chem. C 2015, 119, 28077−28082. (52) Trozzi, F.; Marforio, T. D.; Bottoni, A.; Zerbetto, F.; Calvaresi, M. Engineering the Fullerene-Protein Interface by Computational Design: The Sum Is More than Its Parts. Isr. J. Chem. 2017, 57, 547− 552. (53) Ahmed, L.; Rasulev, B.; Kar, S.; Krupa, P.; Mozolewska, M. A.; Leszczynski, J. Inhibitors or Toxins? Large Library Target-Specific Screening of Fullerene-Based Nanoparticles for Drug Design Purpose. Nanoscale 2017, 9, 10263−10276. (54) Soldà, A.; Cantelli, A.; Di Giosia, M.; Montalti, M.; Zerbetto, F.; Rapino, S.; Calvaresi, M. C60@lysozyme: A New Photosensitizing Agent for Photodynamic Therapy. J. Mater. Chem. B 2017, 5, 6608− 6615. (55) Berger, C. S.; Marks, J. W.; Bolskar, R. D.; Rosenblum, M. G.; Wilson, L. J. Cell Internalization Studies of Gadofullerene-(ZME-018) Immunoconjugates into A375m Melanoma Cells. Transl. Oncol. 2011, 4, 350−354. (56) Li, T.; Murphy, S.; Kiselev, B.; Bakshi, K. S.; Zhang, J.; Eltahir, A.; Zhang, Y.; Chen, Y.; Zhu, J.; Davis, R. M.; Madsen, L. A.; Morris, J. R.; Karolyi, D. R.; LaConte, S. M.; Sheng, Z.; Dorn, H. C. A New Interleukin-13 Amino-Coated Gadolinium Metallofullerene Nanoparticle for Targeted MRI Detection of Glioblastoma Tumor Cells. J. Am. Chem. Soc. 2015, 137, 7881−7888. (57) Han, Z.; Wu, X.; Roelle, S.; Chen, C.; Schiemann, W. P.; Lu, Z.R. Targeted Gadofullerene for Sensitive Magnetic Resonance Imaging and Risk-Stratification of Breast Cancer. Nat. Commun. 2017, 8, 692. (58) Caravan, P. Protein-Targeted Gadolinium-Based Magnetic Resonance Imaging (MRI) Contrast Agents: Design and Mechanism of Action. Acc. Chem. Res. 2009, 42, 851−862. (59) Kang, S.; Huynh, T.; Zhou, R. Non-Destructive Inhibition of Metallofullerenol Gd@C82(OH)22 on WW Domain: Implication on Signal Transduction Pathway. Sci. Rep. 2012, 2, No. 957. (60) Kang, S.; Zhou, G.; Yang, P.; Liu, Y.; Sun, B.; Huynh, T.; Meng, H.; Zhao, L.; Xing, G.; Chen, C.; Zhao, Y.; Zhou, R. Molecular 13788

DOI: 10.1021/acsomega.8b01888 ACS Omega 2018, 3, 13782−13789

ACS Omega

Article

Mechanism of Pancreatic Tumor Metastasis Inhibition by Gd@ C82(OH)22 and Its Implication for de Novo Design of Nanomedicine. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 15431−15436. (61) Pan, Y.; Wang, L.; Kang, S.; Lu, Y.; Yang, Z.; Huynh, T.; Chen, C.; Zhou, R.; Guo, M.; Zhao, Y. Gd−Metallofullerenol Nanomaterial Suppresses Pancreatic Cancer Metastasis by Inhibiting the Interaction of Histone Deacetylase 1 and Metastasis-Associated Protein 1. ACS Nano 2015, 9, 6826−6836. (62) Lu, J.; Mei, W. N.; Gao, Y.; Zeng, X.; Jing, M.; Li, G.; Sabirianov, R.; Gao, Z.; You, L.; Xu, J.; Yu, D.; Ye, H. Structural and Electronic Properties of Gd@C60: All-Electron Relativistic TotalEnergy Study. Chem. Phys. Lett. 2006, 425, 82−84. (63) Wang, C.; Greene, D.; Xiao, L.; Qi, R.; Luo, R. Recent Developments and Applications of the MMPBSA Method. Front. Mol. Biosci. 2018, 4, 87. (64) Guo, J.; Yao, X.; Ning, L.; Wang, Q.; Liu, H. The Adsorption Mechanism and Induced Conformational Changes of Three Typical Proteins with Different Secondary Structural Features on Graphene. RSC Adv. 2014, 4, 9953−9962. (65) Zhang, L.; Xiao, X.; Yuan, Y.; Guo, Y.; Li, M.; Pu, X. Probing Immobilization Mechanism of Alpha-Chymotrypsin onto Carbon Nanotube in Organic Media by Molecular Dynamics Simulation. Sci. Rep. 2015, 5, No. 9297. (66) Zhang, L.; Li, Y.; Yuan, Y.; Jiang, Y.; Guo, Y.; Li, M.; Pu, X. Molecular Mechanism of Carbon Nanotube to Activate Subtilisin Carlsberg in Polar and Non-Polar Organic Media. Sci. Rep. 2016, 6, No. 36838. (67) Zhang, Y.-J.; Ding, J.-N.; Zhong, H.; Sun, C.-P.; Han, J.-G. Molecular Dynamics Exploration of the Binding Mechanism and Properties of Single-Walled Carbon Nanotube to WT and Mutant VP35 FBP Region of Ebola Virus. J. Biol. Phys. 2017, 43, 149−165. (68) Zhang, L.; Yuan, Y.; Ren, T.; Guo, Y.; Li, C.; Pu, X. Shining Light on Molecular Mechanism for Odor-Selectivity of CNTImmobilized Olfactory Receptor. Sci. Rep. 2018, 8, No. 7824. (69) Calvaresi, M.; Zerbetto, F. The Devil and Holy Water: Protein and Carbon Nanotube Hybrids. Acc. Chem. Res. 2013, 46, 2454− 2463. (70) Marchesan, S.; Prato, M. Under the Lens: Carbon Nanotube and Protein Interaction at the Nanoscale. Chem. Commun. 2015, 51, 4347−4359. (71) De Leo, F.; Magistrato, A.; Bonifazi, D. Interfacing Proteins with Graphitic Nanomaterials: From Spontaneous Attraction to Tailored Assemblies. Chem. Soc. Rev. 2015, 44, 6916−6953. (72) Oliveira, S. F.; Bisker, G.; Bakh, N. A.; Gibbs, S. L.; Landry, M. P.; Strano, M. S. Protein Functionalized Carbon Nanomaterials for Biomedical Applications. Carbon 2015, 95, 767−779. (73) Antonucci, A.; Kupis-Rozmysłowicz, J.; Boghossian, A. A. Noncovalent Protein and Peptide Functionalization of Single-Walled Carbon Nanotubes for Biodelivery and Optical Sensing Applications. ACS Appl. Mater. Interfaces 2017, 9, 11321−11331. (74) Iwashita, K.; Shiraki, K.; Ishii, R.; Tanaka, T.; Hirano, A. Liquid Chromatographic Analysis of the Interaction between Amino Acids and Aromatic Surfaces Using Single-Wall Carbon Nanotubes. Langmuir 2015, 31, 8923−8929. (75) Hirano, A.; Kameda, T.; Wada, M.; Tanaka, T.; Kataura, H. Carbon Nanotubes Facilitate Oxidation of Cysteine Residues of Proteins. J. Phys. Chem. Lett. 2017, 8, 5216−5221. (76) Hirano, A.; Kameda, T.; Sakuraba, S.; Wada, M.; Tanaka, T.; Kataura, H. Disulfide Bond Formation of Thiols by Using Carbon Nanotubes. Nanoscale 2017, 9, 5389−5393. (77) Calvaresi, M.; Hoefinger, S.; Zerbetto, F. Probing the Structure of Lysozyme-Carbon-Nanotube Hybrids with Molecular Dynamics. Chem. - Eur. J. 2012, 18, 4308−4313. (78) Hirano, A.; Tanaka, T.; Kataura, H.; Kameda, T. Arginine Side Chains as a Dispersant for Individual Single-Wall Carbon Nanotubes. Chem. - Eur. J. 2014, 20, 4922−4930. (79) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H.

P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision D.1; Gaussian Inc: Wallingford CT, 2009. (80) Dai, X.; Gao, Y.; Xin, M.; Wang, Z.; Zhou, R. The Ground State and Electronic Structure of Gd@C82: A Systematic Theoretical Investigation of First Principle Density Functionals. J. Chem. Phys. 2014, 141, No. 244306. (81) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (82) Cundari, T. R.; Stevens, W. J. Effective Core Potential Methods for the Lanthanides. J. Chem. Phys. 1993, 98, 5555−5565. (83) Dimelow, R. J.; Burton, N. A.; Hillier, I. H. The Dynamics of Water Exchange in Gadolinium DOTA Complexes Studied by Transition Path Sampling and Potential of Mean Force Methods. Phys. Chem. Chem. Phys. 2007, 9, 1318−1323. (84) Calvaresi, M.; Zerbetto, F. In Silico Carborane Docking to Proteins and Potential Drug Targets. J. Chem. Inf. Model. 2011, 51, 1882−1896. (85) Schneidman-Duhovny, D.; Inbar, Y.; Polak, V.; Shatsky, M.; Halperin, I.; Benyamini, H.; Barzilai, A.; Dror, O.; Haspel, N.; Nussinov, R.; Wolfson, H. J. Taking Geometry to Its Edge: Fast Unbound Rigid (and Hinge-Bent) Docking. Proteins 2003, 52, 107− 112. (86) Andrusier, N.; Nussinov, R.; Wolfson, H. J. FireDock: Fast Interaction Refinement in Molecular Docking. Proteins 2007, 69, 139−159. (87) Kingsford, C. L.; Chazelle, B.; Singh, M. Solving and Analyzing Side-Chain Positioning Problems Using Linear and Integer Programming. Bioinformatics 2005, 21, 1028−1039. (88) Case, D. A.; Darden, T. A.; Cheatham, T. E., III; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Walker, R. C.; Zhang, W.; Merz, K. M.; Roberts, B.; Hayik, S.; Roitberg, A.; Seabra, G.; Swails, J.; Goetz, A. W.; Kolossváry, I.; Wong, K. F.; Paesani, F.; Vanicek, J.; Wolf, R. M.; Liu, J.; Wu, X.; Brozell, S. R.; Steinbrecher, T.; Gohlke, H.; Cai, Q.; Ye, X.; Wang, J.; Hsieh, M. J.; Cui, G.; Roe, D. R.; Mathews, D. H.; Seetin, M. G.; Salomon-Ferrer, R.; Sagui, C.; Babin, V.; Luchko, T.; Gusarov, S.; Kovalenko, A.; Kollman, P. A. AMBER 12; University of California: San Francisco, 2012. (89) Tsui, V.; Case, D. A. Theory and Applications of the Generalized Born Solvation Model in Macromolecular Simulations. Biopolymers 2000, 56, 275−291.

13789

DOI: 10.1021/acsomega.8b01888 ACS Omega 2018, 3, 13782−13789