Next Article in Journal
The History of Nanoscience and Nanotechnology: From Chemical–Physical Applications to Nanomedicine
Previous Article in Journal
Enhanced Heavy Metal Removal from Acid Mine Drainage Wastewater Using Double-Oxidized Multiwalled Carbon Nanotubes
Previous Article in Special Issue
Gold Nanoparticles as Boron Carriers for Boron Neutron Capture Therapy: Synthesis, Radiolabelling and In Vivo Evaluation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Neutral Rhenadicarbaboranes with Re(CO)2(NO) Vertices: A Theoretical Study of Building Blocks for Rhenacarborane-Based Drug Delivery Agents

by
Amr A. A. Attia
1,
Alexandru Lupan
1,*,
Radu Silaghi-Dumitrescu
1 and
R. Bruce King
2,*
1
Faculty of Chemistry and Chemical Engineering, Babeș-Bolyai University, Cluj-Napoca 400084, Romania
2
Department of Chemistry, University of Georgia, Athens, GA 30602, USA
*
Authors to whom correspondence should be addressed.
Molecules 2020, 25(1), 110; https://doi.org/10.3390/molecules25010110
Submission received: 6 November 2019 / Revised: 18 December 2019 / Accepted: 20 December 2019 / Published: 27 December 2019
(This article belongs to the Special Issue Advances in Materials Derived from Polyhedral Boron Clusters)

Abstract

:
The rhenadicarbaborane carbonyl nitrosyls (C2Bn−3Hn−1)Re(CO)2(NO), (n = 8 to 12), of interest in drug delivery agents based on the experimentally known C2B9H11Re(CO)2(NO) and related species, have been investigated by density functional theory. The lowest energy structures of these rhenadicarbaboranes are all found to have central ReC2Bn−3 most spherical closo deltahedra in accord with their 2n + 2 Wadean skeletal electrons. Carbon atoms are found to be located preferentially at degree 4 vertices in such structures. Furthermore, rhenium atoms are preferentially located at a highest degree vertex, typically a vertex of degree 5. Only for the 9-vertex C2B6H8Re(CO)2(NO) system are alternative isocloso deltahedral isomers found within ~8 kcal/mol of the lowest energy closo isomer. Such 9-vertex isocloso structures provide a degree 6 vertex for the rhenium atom flanked by degree 4 vertices for each carbon atom.

Graphical Abstract

1. Introduction

Carboranes in general are accepted as synthons not unlike organic ones insofar as biological and medical applications are concerned [1]. The very common icosahedral C2B10 synthon is thus regarded as similar to rotating phenyl groups. This similarity is seen in terms of steric requirements, polarity/hydrophobicity, and availability of regioselective functionalization with substituents at the carbon atoms [1]. Such regioselective functionalization has been shown to work with all classes of bioactive substances, including aminoacids, lipids, and nucleosides, as well as incorporation into dendrimers, liposomes, and other biocompatible methods of nanoencapsulation [1]. In this connection carboranes also provide two specific advantages: the presence of boron, which allows for radiotherapeutical techniques such as boron neutron capture therapy (BNCT), and the availability of viable syntheses leading to endo-substitution with transition metal synthons to yield metallacarboranes [1,2,3,4,5,6,7,8,9].
With such potential, carboranes have been shown to act as agonists or, depending on the substitution, antagonists, of the estrogen receptor protein [4]. Thus the steric and hydrophobic similarity/compatibility of the carborane unit with the non-aromatic part of the estrogen allows it to compete efficiently with the latter, either as a substitute or as an inhibitor [1]. Carboranes as hydrophobic units can be incorporated into a host of other proteins. These include the androgen receptors (e.g., with testosterone as target, leading to carborane-containing derivatives shown to outperform currently used drugs against prostate cancer) [10], retinoic acid receptors (with successful tests performed on human promyelelocytic leukemia HL-60 cells—thus again with anti-cancer potential) [11]. Other such proteins include transthyretin (a target of non-steroidal anti-inflammatory drugs), NSAIDS, where carborane derivatives have the unique advantage of not showing concomitant affinity for cyclooxygenase enzymes, COX, and hence displaying reduced side-effects as anti-inflammatory agents compared to typical NSAIDS. Interestingly, the lack of inhibition in cyclooxygenase also arises from the analogy of carboranyl with phenyl. However, in this case COX inhibition would require a non-rotating phenyl ring) [1]. Other such proteins include HIV protease (in this case with ionic carborane derivatives but still with a key role for hydrophobic interactions) [12,13], α-thrombin (acting as anticoagulants) [11], as well as a range of derivatives aimed at accumulating boron in tumor cells for BNCT (e.g., with nucleosides). A more recent application involves BNCS (boron neutron capture synovectomy, specifically for removal of synovial tissue) [14,15], or coupled with cytotoxic metals such as platinum or tin, or coupled with light-sensitive moieties for use in photodynamic therapy (PDT) [1]. The range of metals of the metallacarboranes exhibiting biological reactivity as described above includes Fe, Co, V, Ta, Mo, Nb, Tc, Re [1]. A particular set of applications is the one regarding radioimaging and radiotherapy. For imaging, in addition to halogen radioactive isotopes, technetium is the most widely used [1]. Rhenium (and hence rhenacarboranes) has been proposed as a convenient substitute for technetium for preliminary laboratory and in vitro studies, since the two metals are reasonably similar in properties insofar their metallacarboranes are concerned. In addition, rhenium is much more readily available and non-radioactive [1,10,11].
Rhenium has also been shown to be useful independently of technetium, such as in the iodine-radioactive 3-NO-3,3-κ2-(2,2′-N2C10H6(Me){(CH2)7131I}-4,4′)-closo-3,1,2-ReC2B9H11 rhenacarborane and in subsequently reported members of its family, [3,3-(CO)2-3-NO-closo-Re(8-O(CH2)2O(CH2)2NH3-3,1,2-C2B9H10)]BF4 and [3,3-(CO)2-3-NO-closo-Re(8-O(CH2)2O(CH2)2OH-3,1,2-C2B9H10)] with a range of other compounds also available synthetically [1]. Such molecules have the remarkable property of passing the blood-brain barrier with the hydrophobicity of the carborane and the charge-neutral character of the nitrosyl/carbonyl-metal moiety being essential towards achieving this goal. Such observations suggest clear potential for therapeutic agents directed within the central nervous system either by delivering radioactive material via the substituents of the carborane system, or by delivering other conjugated therapeutic agents which would not otherwise pass the blood-brain barrier, such as peptides [1,7]. Also, when rhenacarboranes are compared to their technetium analogs, they complement their lack of radioactivity with a useful degree of luminescence with a potential for in vitro and in vivo imaging. The latter, correlated with the distinct affinity for biological targets such as the estrogen receptor, has fueled further synthetic interest into rhenacarboranes and into their biologically-relevant reactivity, structure and stability [1,2,3,4,5,6,7,8,9].
The chemistry of such rhenacarboranes dates back to the 1965 synthesis of the icosahedral dicarbaborane rhenium carbonyl monoanion [3,1,2-C2B9H11Re(CO)3] as its cesium salt [16,17]. Replacement of one of the carbonyl groups in [3,1,2-C2B9H11Re(CO)3] with NO+ gives the neutral 3,1,2-C2B9H11Re(CO)2(NO) originally synthesized by Stone and co-workers [18] but then later used by Jelliss and coworkers as a building block for their work on rhenacarborane-based drug delivery agents [1,7]. Related neutral rhenacarborane derivatives include the tricarbaborane derivatives (C3Bn−4Hn−1)Re(CO)3, which are direct analogues of the stable (η5-C5H5)Re(CO)3. However, suitable precursors for rhenatricarbaborane derivatives are less accessible synthetically than those for rhenadicarbaboranes since dicarbaborane precursors are readily accessible from boranes and alkynes. The chemical robustness of certain rhenacarboranes has additionally made them of interest as possible vehicles for drug delivery in therapeutic and diagnostic applications [19]. Thus suitably designed kinetically inert rhenacarborane polyhedra can survive reaction conditions required to introduce external functionalities for optimal introduction into biological systems. In addition, such rhenacarboranes can survive metabolic degradation.
The central polyhedra in the three types of rhenacarboranes, namely [(C2Bn−3Hn−1)Re(CO)3], (C2Bn−3Hn−1)Re(CO)2(NO), and (C3Bn−4Hn−1)Re(CO)3, all have 2n + 2 skeletal electrons, where BH, CH, Re(CO)3, and Re(CO)2(NO) vertices contribute 2, 3, 1, and 2 skeletal electrons, respectively, assuming contributions of three internal orbitals from each vertex atom towards the skeletal bonding. Such rhenacarboranes therefore might be expected to exhibit most spherical closo deltahedral structures (Figure 1) in accord with the Wade-Mingos rules [20,21,22]. The closo deltahedra for the 8- through 12-vertex systems have only degree 4 and 5 vertices except for the 11-vertex system, which necessarily has one degree 6 vertex. However, density functional theory studies [23] show that the asymmetry in the C3Bn−4Re polyhedra leads to deviation from the most spherical deltahedra to give lowest energy structures for the 8- and 10-vertex systems with two degree 6 vertices. We now report similar density theory functional studies on the neutral (C2Bn−3Hn−1)Re(CO)2(NO) systems (n = 9 to 12), which are the basis for the rhenacarborane drug delivery systems.
Alternative deltahedra to be considered for the 9- and 10-vertex rhenadicarboranes are the isocloso deltahedra, which, unlike the corresponding closo deltahedra, provide a degree 6 vertex for the rhenium atom (Figure 2) [24,25,26]. Conversion of a closo deltahedron to an isocloso deltahedron with the same number of vertices involves a diamond-square-diamond rearrangement converting a pair of degree 5 vertices to a degree 6 and a degree 4 vertex. Normally isocloso deltahedral metallaborane structures are found in systems with 2n skeletal electrons rather than 2n + 2 skeletal electrons. However, the closoisocloso metallaborane conversion for the 9- and 10-vertex rhenacarboranes can provide a degree 6 vertex for the rhenium atom and an additional more favorable degree 4 rather than a degree 5 vertex for a carbon atom. For the 10-vertex rhenatricarbaborane C3B6H9Re(CO)3 the lowest energy structures are found to be isocloso rather than closo deltahedra (compare Figure 1 and Figure 2) thereby providing a favorable degree 4 vertex for each of the three carbon atoms in addition to a degree 6 vertex for the rhenium atom.

2. Results and Discussion

2.1. Eight-Vertex C2B5H7Re(CO)2(NO) Structures

Five 8-vertex C2B5H7Re(CO)2(NO) structures were found within 26 kcal/mol of the lowest energy structure B5C2Re-1 (Figure 3 and Table 1). All of these five structures were found to have the expected central ReC2B5 bisdisphenoid, which is the most spherical closo deltahedron consistent with the 2n + 2 Wadean skeletal electrons in this system. The four lowest-energy structures have the rhenium atom located at a degree 5 vertex and both carbon atoms located at non-adjacent degree 4 vertices separated by a single boron atom. This corresponds to C…C distances in the range 2.6 to 2.7 Å. The lowest energy structure B5C2Re-1 as well as B5C2Re-2, lying 4.3 kcal/mol in energy above B5C2Re-1, have one Re–C edge and differ by rotation of the Re(CO)2(NO) moiety. The next higher energy C2B5H7Re(CO)2(NO) structures, namely B5C2Re-3 and B5C2Re-4, lying 7.2 and 11.0 kcal/mol, respectively, in energy above B5C2Re-1, have two Re–C edges and again differ mainly by rotation of the Re(CO)2(NO) moiety. Structure B5C2Re-5, lying 13.6 kcal/mol in energy above B5C2Re-1, has the rhenium atom located at a degree 4 vertex and one Re–C edge.

2.2. Nine-Vertex C2B6H8Re(CO)2(NO) Structures

The lowest energy C2B6H8Re(CO)2(NO) structure B6C2Re-1, as well as the next two higher energy structures B6C2Re-2 and B6C2Re-3, lying 2.7 and 3.7 kcal/mol, respectively, in energy above B6C2Re-1, all have a central ReC2B6 tricapped trigonal prism, which is the most spherical closo 9-vertex deltahedron in accord with the Wade–Mingos rules for this 2n + 2 skeletal electron system (Figure 4 and Table 2). All three of these structures have the only possible arrangement of a degree 5 rhenium vertex, and two degree 4 carbon vertices. They differ only in the rotation of the Re(CO)2(NO) unit relative to the ReC2B6 cage. The higher energy C2B6H8Re(CO)2(NO) structure B6C2Re-6, lying 11.6 kcal/mol in energy above C2B6H8Re(CO)2(NO) structure B6C2Re-1, also has a central ReC2B6 tricapped trigonal prism but with one of the carbon atoms located at a less favorable degree 5 vertex.
In addition to these four C2B6H8Re(CO)2(NO) structures with a central ReC2B6 tricapped trigonal prism, three C2B6H8Re(CO)2(NO) structures, namely B6C2Re-4, B6C2Re-5, and B6C2Re-7, lying 8.0, 9.9 and 13.3 kcal/mol in energy above B6C2Re-1, are found with a central ReC2B6 isocloso deltahedron, thereby providing a degree 6 vertex for the rhenium atom (Figure 2 and Figure 4 and Table 2). All three isocloso C2B6H8Re(CO)2(NO) structures have both carbon atoms located at degree 4 vertices necessarily adjacent to the rhenium vertex thus leading to two Re–C edges. Structure B6C2Re-4 with Cs symmetry has the two carbon vertices located on a side of the rectangle of the four degree 4 vertices. However, structures B6C2Re-5 and B6C2Re-7, each with C2 symmetry, have their two carbon vertices located on a diagonal of the rectangle of the four degree 4 vertices. Structures B6C2Re-5 and B6C2Re-7 differ in the rotation of the Re(CO)2(NO) unit relative to the ReC2B6 isocloso deltahedron.

2.3. Ten-Vertex C2B7H9Re(CO)2(NO) Structures

Four 10-vertex C2B7H9Re(CO)2(NO) structures were found within 16 kcal/mol of the lowest energy structure B7C2Re-1 (Figure 5 and Table 3). All of these structures have a central ReC2B7 bicapped tetragonal antiprism, which is the 10-vertex closo deltahedron consistent with the 2n + 2 = 22 skeletal electrons for this system. The lowest energy C2B7H9Re(CO)2(NO) structure B7C2Re-1 as well as the slightly higher energy structure B7C2Re-2, lying only 2.0 kcal/mol in energy above B7C2Re-1, have the only possible arrangement with the two carbon atoms located at the two degree 4 vertices necessarily leading to a single Re–C edge. The antipodal positions of the two carbon vertices in B7C2Re-1 and B7C2Re-2 lead to relative long C…C distances of 3.42 Å. Structures B7C2Re-1 and B7C2Re-2 differ only in the orientation of the Re(CO)2(NO) unit. The two next higher energy C2B7H9Re(CO)2(NO) structures, namely B7C2Re-3 and B7C2Re-4, lying 11.1 and 14.8 kcal/mol in energy, respectively, above B7C2Re-1, have the energetically less desirable feature of one of the carbon atoms located at a degree 5 rather than a degree 4 vertex. In B7C2Re-3, neither carbon vertex is adjacent to the rhenium vertex so there are no Re–C edges in the ReC2B7 deltahedron. However, in B7C2Re-4, the degree 5 carbon vertex is adjacent to the rhenium vertex so that there is an Re–C edge.
The clear energetic preference of the 10-vertex rhenadicarbaborane C2B7H9Re(CO)2(NO) for the closo bicapped tetragonal antiprism structures (Figure 1) contrasts with the previously discovered [23] energetic preference of the 10-vertex rhenatricarbaborane C3B6H9Re(CO)3 for isocloso deltahedral structures (Figure 2). This difference may relate to the number of carbon vertices in the central deltahedron. For the rhenadicarbaboranes C2B7H9Re(CO)2(NO), the closo structure provides the two degree 4 vertices required for the two carbon atoms. However, for the rhenatricarbaboranes C3B6H9Re(CO)3, an isocloso structure is required to provide the three degree 4 vertices to accommodate all three carbon atoms.

2.4. Eleven-Vertex C2B8H10Re(CO)2(NO) Structures

The 11-vertex closo deltahedron (Figure 1) has a single degree 6 vertex and thus also can function as an isocloso metallaborane with a metal atom at the degree 6 vertex. The six lowest energy C2B8H10Re(CO)2(NO) structures are all based on this deltahedron (Figure 6 and Table 4). The lowest energy such structure B8C2Re-1 has the ideal arrangement with the rhenium atom located at the lone degree 6 vertex and the carbon atoms located at the two degree 4 vertices, thereby leading to two Re-C edges. The next highest energy C2B8H10Re(CO)2(NO) structure B8C2Re-2, lying 6.6 kcal/mol in energy above B8C2Re-1, has the rhenium atom still located at the degree 6 vertex but one of the carbon atoms has moved from a degree 4 vertex to a degree 5 vertex not adjacent to the other carbon atom. Structure B8C2Re-4, lying 10.7 kcal/mol in energy above B8C2Re-1, also has the rhenium atom located at the degree 6 vertex, one carbon atom located at a degree 4 vertex, and the other carbon atom located at a degree 5 vertex different from the degree 5 carbon vertex in B8C2Re-2.
The three remaining C2B8H10Re(CO)2(NO) structures, namely B8C2Re-3, B8C2Re-5, and B8C2Re-6, lying 7.1, 11.3, and 12.3 kcal/mol, respectively, in energy above B8C2Re-1, each have the two carbon atoms located at their two degree 4 vertices and the rhenium atom located at a degree 5 vertex (Figure 6 and Table 4). These three structures differ in the location of the degree 5 rhenium vertex relative to the degree 6 and degree 4 vertices of the underlying 11-vertex closo deltahedron. Thus, in B8C2Re-5, the rhenium atom is located at the vertex furthest from the degree 6 vertex leading to an Re–B (deg 6) distance of 3.54 Å. In B8C2Re-6, the rhenium atom is located at a degree 5 vertex adjacent to the degree 6 vertex with a Re–B (deg 6) edge of length 2.65 Å. However, in B8C2Re-3, the rhenium atom is located at a degree 5 vertex adjacent to a degree 4 vertex but not adjacent to the degree 6 vertex leading to an an Re–B (deg 6) distance of 3.40 Å.

2.5. Twelve-Vertex C2B9H11Re(CO)2(NO) Structures

The five lowest energy 12-vertex C2B9H11Re(CO)2(NO) structures are all based on the regular icosahedron with all degree 5 vertices consistent with their 2n + 2 skeletal electrons for this closo deltahedron (Figure 1 and Figure 7, Table 5). None of these five icosahedral C2B9H11Re(CO)2(NO) have a C–C edge consistent with the pattern of the smaller (C2Bn−3Hn−1)Re(CO)2(NO) structures discussed above. The lowest energy C2B9H11Re(CO)2(NO) structure B9C2Re-1 is the unique structure not only without a C-C edge but also without any Re–C edges. The nest lowest energy C2B9H11Re(CO)2(NO) structure B9C2Re-2, lying only 1.8 kcal/mol in energy above B9C2Re-1, is the unique icosahedral structure with the carbon atoms in antipodal positions leading to a C…C distance of 3.05 Å. The next three C2B9H11Re(CO)2(NO) structures, namely B9C2Re-3, B9C2Re-4, and B9C2Re-5, lying 2.6, 4.6, and 6.1 kcal/mol in energy above B9C2Re-1, have the pair of carbon atoms located in non-adjacent non-antipodal positions separated by a single boron vertex in one direction leading to C–C distances of ~2.6 Å. Structures B9C2Re-3 and B9C2Re-4 have one Re–C edge whereas, B9C2Re-5 is the unique possible icosahedral structure with two Re–C edges.

3. Theoretical Methods

The rhenadicarbaborane structures investigated in this study are derived from the borane dianions BnHn2− by substituting a BH vertex with an Re(CO)2(NO) unit followed by the replacement of two additional boron atoms with two carbon atoms. A total of 48 polyhedral frameworks ranging from 8 to 12 vertices were generated in this way leading to 2426 initial structures of the type C2Bn−3Hn−1Re(CO)2(NO) (n = 8 to 12) (see Supporting Information).
Geometry optimizations of these initial structures were performed using the B3LYP DFT functional coupled with the SDD (Stuttgart Dresden ECP plus DZ) basis set for rhenium and the double zeta 6-31G(d) basis set for the lighter atoms as implemented in the Gaussian09 suite of programs [27]. All optimized structures were characterized by harmonic vibrational frequencies. Saddle point structures with imaginary vibrational frequencies were reoptimized by following the normal modes in order to obtain genuine minima. The energetically most stable isomers were further optimized by employing the M06L DFT functional and the 6-311G(d,p)//SDD basis sets. All of the resulting structures were found to have substantial HOMO-LUMO gaps ranging from 2.2 to 3.0 eV (see the Supporting Information).
The shorthand notation B(n-3)C2Re-x was assigned to all structures discussed in this work where n is the total number of polyhedral vertices, and x is the energy ranking of the structure on the potential energy surface. Additional information on higher energy structures and connectivities can be viewed in the Supporting Information.

4. Summary

The lowest energy (C2Bn−3Hn−1)Re(CO)2(NO) structures are all found to have central ReC2Bn−3 most spherical closo deltahedra in accord with their 2n + 2 Wadean skeletal electrons. Carbon atoms are preferentially located at degree 4 vertices whereas rhenium atoms are preferentially located at a highest degree vertex, typically a vertex of degree 5. Only for the 9-vertex C2B6H8Re(CO)2(NO) system are alternative isocloso deltahedral isomers found within 8 kcal/mol of the lowest energy closo isomer. Such 9-vertex isocloso structures provide a degree 6 vertex for the rhenium atom with adjacent degree 4 vertices for both carbon atoms.

Supplementary Materials

Table S1A–C. Initial models, distance matrices and energy rankings for the C2B5H7Re(CO)2NO structures. Table S2A–C. Initial models, distance matrices and energy rankings for the C2B6H8Re(CO)2NO structures. Table S3A–C. Initial models, distance matrices and energy rankings for the C2B7H9Re(CO)2NO structures. Table S4A–C. Initial models, distance matrices and energy rankings for the C2B8H10Re(CO)2NO structures. Table S5A–C. Initial models, distance matrices and energy rankings for the C2B9H11Re(CO)2NO structures. Table S6. Orbital energies and HOMO-LUMO gaps. Complete Gaussian09 reference.

Author Contributions

A.L. conceived the project, supervised the calculations, and organized the data; A.A.A.A. performed the calculations; R.S.-D. wrote the portion of the Introduction relating to biological and medical significance of the molecules investigated; and R.B.K. wrote most of the initial draft and prepared the final draft. All authors have read and agreed to the published version of the manuscript.

Funding

Funding from the Romanian Ministry of Education and Research, (Grant PN-III-P4-ID-PCE-2016-0089) is gratefully acknowledged. Computational resources were provided by the high-performance computational facility of the Babeș-Bolyai University (MADECIP, POSCCE, COD SMIS 48801/1862) co-financed by the European Regional Development Fund.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grimes, R.N. Carboranes; Academic Press: Cambridge, MA, USA, 2016; pp. 1053–1082. [Google Scholar]
  2. Armstrong, A.F.; Valliant, J.F. Differences between the macroscopic and tracer level chemistry of rhenium and technetium: Contrasting cage isomerisation behaviour of Re(I) and Tc(I) carborane complexes. Dalton Trans. 2010, 39, 8128–8131. [Google Scholar] [CrossRef] [PubMed]
  3. Armstrong, A.F.; Lebert, J.M.; Brennan, J.D.; Valliant, J.F. Functionalized carborane complexes of the [M(CO)2(NO)]2+ core (M = 99mTc, Re): A new class of organometallic probes for correlated in vitro and in vivo imaging. Organometallics 2009, 28, 2986–2992. [Google Scholar] [CrossRef]
  4. Causey, P.W.; Besanger, T.R.; Valliant, J.F. Synthesis and screening of mono- and di-aryl technetium and rhenium metallocarboranes. A new class of probes for the estrogen receptor. J. Med. Chem. 2008, 51, 2833–2844. [Google Scholar] [CrossRef] [PubMed]
  5. Louie, A.S.; Harrington, L.E.; Valliant, J.F. The preparation and characterization of functionalized carboranes and Re/Tc-metallocarboranes as platforms for developing molecular imaging probes: Structural and cage isomerism studies. Inorg. Chim. Acta 2012, 389, 159–167. [Google Scholar] [CrossRef]
  6. El-Zaria, M.E.; Janzen, N.; Valliant, J.F. Room-temperature synthesis of Re(I) and Tc(I) metallocarboranes. Organometallics 2012, 31, 5940–5949. [Google Scholar] [CrossRef]
  7. Pruitt, D.G.; Baumann, S.M.; Place, G.J.; Oyeamalu, A.N.; Sinn, E.; Jelliss, P.A. Synthesis and functionalization of nitrosyl rhenacarboranes towards their use as drug delivery vehicles. J. Organomet. Chem. 2015, 798, 60–69. [Google Scholar] [CrossRef] [Green Version]
  8. Ma, P.; Smith, T.M.; Zubieta, J.; Spencer, J.T. Synthesis of alkoxy derivatives of the 10-vertex manganadecaborane [nido-6-Mn(CO)3B9H13][NMe4]. Inorg. Chem. Commun. 2014, 46, 223–225. [Google Scholar] [CrossRef]
  9. Genady, A.R.; Tan, J.; El-Zaria, M.E.; Zlitni, A.; Janzen, N.; Valliant, J.F. Synthesis, characterization and radiolabeling of carborane-functionalized tetrazines for use in inverse electron demand Diels–Alder ligation reactions. J. Organomet. Chem. 2015, 791, 204–213. [Google Scholar] [CrossRef]
  10. Stockmann, P.; Gozzi, M.; Kuhnert, R.; Sárosi, M.B.; Hey-Hawkins, E. New keys for old locks: Carborane-containing drugs as platforms for mechanism-based therapies. Chem. Soc. Rev. 2019, 48, 3497–3512. [Google Scholar] [CrossRef] [Green Version]
  11. Issa, F.; Kassiou, M.; Rendina, L.M. Boron in drug discovery: Carboranes as unique pharmacophores in biologically active compounds. Chem. Rev. 2011, 111, 5701–5722. [Google Scholar] [CrossRef]
  12. Fanfrlík, J.; Hnyk, D.; Lepsík, M.; Hobza, P. Interaction of heteroboranes with biomolecules. Part 2. The effect of various metal vertices and exo-substitutions. Phys. Chem. Chem. Phys. 2007, 9, 2085–2093. [Google Scholar] [CrossRef] [PubMed]
  13. Cígler, P.; Kozísek, M.; Rezácová, P.; Brynda, J.; Otwinowski, Z.; Pokorná, J.; Plešek, J.; Grüner, B.; Dolecková-Maresová, L.; Mása, M.; et al. From nonpeptide toward noncarbon protease inhibitors: Metallacarboranes as specific and potent inhibitors of HIV protease. Proc. Natl. Acad. Sci. USA 2005, 102, 15394–15399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Valliant, J.F.; Schaffer, P. A new approach for the synthesis of isonitrile carborane derivatives. Ligands for metal based boron neutron capture therapy (BNCT) and boron neutron capture synovectomy (BNCS) agents. J. Inorg. Biochem. 2001, 85, 43–51. [Google Scholar] [CrossRef]
  15. Valliant, J.F.; Sogbein, O.O.; Morel, P.; Schaffer, P.; Guenther, K.J.; Bain, A.D. Synthesis, NMR, and X-ray crystallographic analysis of C-hydrazino-C-carboxycarboranes: Versatile ligands for the preparation of BNCT and BNCS agents and (99m) Tc radiopharmaceuticals. Inorg. Chem. 2002, 41, 2731–2737. [Google Scholar] [CrossRef] [PubMed]
  16. Hawthorne, M.F.; Andrews, T.D. Carbametallic boron hydride derivatives. II. apparent analogs of π-C5H5Mn(CO)3 and π-C5H5Re(CO)3. J. Am. Chem. Soc. 1965, 87, 2496. [Google Scholar] [CrossRef]
  17. Andrews, T.D.; Hawthorne, M.F.; Howe, D.V.; Pilling, R.L.; Pitts, D.; Reintjes, M.; Warren, L.F.; Wegner, P.A.; Young, D.C. π-Dicarbollyl derivatives of the transition metals. Metallocene analogs. J. Am. Chem. Soc. 1968, 90, 879–896. [Google Scholar]
  18. Ellis, D.D.; Jelliss, P.A.; Stone, F.G.A. Rhenacarborane complexes with nitrosyl and alkylidene ligands. Chem. Commun. 1999, 23, 2385–2386. [Google Scholar] [CrossRef]
  19. Gozzi, M.; Schwarze, B.; Hey-Hawkins, E. Half- and mixed-sandwich metallacarboranes for potential applications in medicine. Pure Appl. Chem. 2019, 91, 563–573. [Google Scholar] [CrossRef]
  20. Wade, K. The structural significance of the number of skeletal bonding electron-pairs in carboranes, the higher boranes and borane anions, and various transition-metal carbonyl cluster compounds. Chem. Commun. 1971, 15, 792–793. [Google Scholar] [CrossRef]
  21. Mingos, D.M.P. A general theory for cluster and ring compounds of the main group and transition elements. Nat. Phys. Sci. 1972, 236, 99–102. [Google Scholar] [CrossRef]
  22. Mingos, D.M.P. Polyhedral skeletal electron pair approach. Acc. Chem. Res. 1984, 17, 311–319. [Google Scholar] [CrossRef]
  23. Attia, A.A.A.; Lupan, A.; King, R.B. Deviations from the most spherical deltahedra in rhenatricarbaboranes having 2n + 2 wadean skeletal electrons. Inorg. Chem. 2017, 56, 15015–15025. [Google Scholar] [CrossRef] [PubMed]
  24. Bould, J.; Kennedy, J.D.; Thornton-Pett, M. Ten-vertex metallaborane chemistry. Aspects of the iridadecaborane closoisonidoisocloso structural continuum. J. Chem. Soc. Dalton Trans. 1992, 4, 563–576. [Google Scholar] [CrossRef]
  25. Kennedy, J.D. The Borane-Carborane-Carbocation Continuum; Casanova, J., Ed.; Wiley: New York, NY, USA, 1998; Chapter 3; pp. 85–116. [Google Scholar]
  26. Štibr, B.; Kennedy, J.D.; Drdáková, E.; Thornton-Pett, M. Nine-vertex polyhedral iridamonocarbaborane chemistry. Products of thermolysis of [(CO)(PPh3)2IrCB7H8] and emerging alternative cluster-geometry patterns. J. Chem. Soc. Dalton 1994, 2, 229–236. [Google Scholar] [CrossRef]
  27. Gaussian, Inc. Gaussian09 (Revision E.01); Gaussian, Inc.: Wallingford, CT, USA, 2009; The Complete Reference Is Given in the Supporting Information. [Google Scholar]
Figure 1. The most spherical (closo) deltahedra having from 8 to 12 vertices indicating their vertex degrees. Vertices of degree 4, 5, and 6 are also indicated in red, black, and green, respectively, in Figure 1 and Figure 2.
Figure 1. The most spherical (closo) deltahedra having from 8 to 12 vertices indicating their vertex degrees. Vertices of degree 4, 5, and 6 are also indicated in red, black, and green, respectively, in Figure 1 and Figure 2.
Molecules 25 00110 g001
Figure 2. The isocloso deltahedra for the 9- and 10-vertex systems.
Figure 2. The isocloso deltahedra for the 9- and 10-vertex systems.
Molecules 25 00110 g002
Figure 3. The five optimized lowest energy 8-vertex C2B5H7Re(CO)2(NO) structures.
Figure 3. The five optimized lowest energy 8-vertex C2B5H7Re(CO)2(NO) structures.
Molecules 25 00110 g003
Figure 4. The seven optimized lowest energy 9-vertex C2B6H8Re(CO)2(NO) structures.
Figure 4. The seven optimized lowest energy 9-vertex C2B6H8Re(CO)2(NO) structures.
Molecules 25 00110 g004
Figure 5. The four optimized lowest energy 10-vertex C2B7H9Re(CO)2(NO) structures.
Figure 5. The four optimized lowest energy 10-vertex C2B7H9Re(CO)2(NO) structures.
Molecules 25 00110 g005
Figure 6. The six optimized lowest energy 11-vertex C2B8H10Re(CO)2(NO) structures.
Figure 6. The six optimized lowest energy 11-vertex C2B8H10Re(CO)2(NO) structures.
Molecules 25 00110 g006
Figure 7. The five optimized lowest energy 12-vertex C2B9H11Re(CO)2(NO) structures.
Figure 7. The five optimized lowest energy 12-vertex C2B9H11Re(CO)2(NO) structures.
Molecules 25 00110 g007
Table 1. The five optimized 8-vertex C2B5H7Re(CO)2(NO) structures within 26 kcal/mol of the lowest energy structure.
Table 1. The five optimized 8-vertex C2B5H7Re(CO)2(NO) structures within 26 kcal/mol of the lowest energy structure.
Vertex DegreesRe–CC–C
StructureEReCEdgesDistance, ÅComments
B5C2Re-10.054,412.60bisdisphenoid
B5C2Re-24.354,412.59bisdisphenoid
B5C2Re-37.254,422.70bisdisphenoid
B5C2Re-411.054,422.70bisdisphenoid
B5C2Re-513.644,412.59bisdisphenoid
Table 2. The seven optimized 9-vertex C2B6H8Re(CO)2(NO) structures within 17 kcal/mol of the lowest energy structure.
Table 2. The seven optimized 9-vertex C2B6H8Re(CO)2(NO) structures within 17 kcal/mol of the lowest energy structure.
Structure Vertex DegreesRe–CC–C
(Symmetry)EReCEdgesDistance, ÅComments
B6C2Re-1 (C1)0.054.412.56tricap trig prism
B6C2Re-2 (C1)2.754,412.57tricap trig prism
B6C2Re-3 (C1)3.754,412.55tricap trig prism
B6C2Re-4 (Cs)8.064,422.77(m)9-vertex isocloso
B6C2Re-5 (C2)9.964,423.17(p)9-vertex isocloso
B6C2Re-6 (Cs)11.655,412.61tricap trig prism
B6C2Re-7 (C2)13.364,423.18(p)9-vertex isocloso
Table 3. The four optimized 10-vertex C2B7H9Re(CO)2(NO) structures within 16 kcal/mol of the lowest energy structure.
Table 3. The four optimized 10-vertex C2B7H9Re(CO)2(NO) structures within 16 kcal/mol of the lowest energy structure.
Structure Vertex DegreesRe–CC–C
(Symmetry)EReCEdgesDistance, ÅComments
B7C2Re-1 (Cs)0.054,413.42Bicap tetrag antipr
B7C2Re-2 (Cs)2.054,413.42Bicap tetrag antipr
B7C2Re-3 (Cs)11.155,402.60Bicap tetrag antipr
B7C2Re-4 (C1)14.855,412.64Bicap tetrag antipr
Table 4. The six optimized 11-vertex C2B8H10Re(CO)2(NO) structures within 15 kcal/mol of the lowest energy structure.
Table 4. The six optimized 11-vertex C2B8H10Re(CO)2(NO) structures within 15 kcal/mol of the lowest energy structure.
Structure Vertex DegreesRe–CC–C
(Symmetry)EReCEdgesDistance, ÅComments
B8C2Re-1 (C2v)0.064.423.3611-vertex closo
B8C2Re-2 (C1)6.665.412.6011-vertex closo
B8C2Re-3 (Cs)7.154,413.0411-vertex closo
B8C2Re-4 (C1)10.765,412.8511-vertex closo
B8C2Re-5 (Cs)11.354,403.0111-vertex closo
B8C2Re-6 (C1)12.354,413.0211-vertex closo
Table 5. The five optimized 12-vertex C2B9H11Re(CO)2(NO) structures within 15 kcal/mol of the lowest energy structure. All of these structures have central C2B9Re icosahedra with exclusively degree 5 vertices.
Table 5. The five optimized 12-vertex C2B9H11Re(CO)2(NO) structures within 15 kcal/mol of the lowest energy structure. All of these structures have central C2B9Re icosahedra with exclusively degree 5 vertices.
Re–CC–C
Structure (Symmetry)EEdgesDistance, Å
B9C2Re-1 (C1)0.002.58
B9C2Re-2 (C1)1.813.05
B9C2Re-3 (C1)2.612.60
B9C2Re-4 (C1)4.612.60
B9C2Re-5 (C1)6.122.67

Share and Cite

MDPI and ACS Style

Attia, A.A.A.; Lupan, A.; Silaghi-Dumitrescu, R.; King, R.B. Neutral Rhenadicarbaboranes with Re(CO)2(NO) Vertices: A Theoretical Study of Building Blocks for Rhenacarborane-Based Drug Delivery Agents. Molecules 2020, 25, 110. https://doi.org/10.3390/molecules25010110

AMA Style

Attia AAA, Lupan A, Silaghi-Dumitrescu R, King RB. Neutral Rhenadicarbaboranes with Re(CO)2(NO) Vertices: A Theoretical Study of Building Blocks for Rhenacarborane-Based Drug Delivery Agents. Molecules. 2020; 25(1):110. https://doi.org/10.3390/molecules25010110

Chicago/Turabian Style

Attia, Amr A. A., Alexandru Lupan, Radu Silaghi-Dumitrescu, and R. Bruce King. 2020. "Neutral Rhenadicarbaboranes with Re(CO)2(NO) Vertices: A Theoretical Study of Building Blocks for Rhenacarborane-Based Drug Delivery Agents" Molecules 25, no. 1: 110. https://doi.org/10.3390/molecules25010110

Article Metrics

Back to TopTop