Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Tuning the selectivity of P4 reduction at alkaline-earth metal centres

Stefan Thum , Oliver P. E. Townrow , Jens Langer and Sjoerd Harder *
Inorganic and Organometallic Chemistry, Friedrich-Alexander-Universität Erlangen-Nürnberg, Egerlandstraße 1, 91058 Erlangen, Germany. E-mail: sjoerd.harder@fau.de

Received 16th December 2024 , Accepted 5th February 2025

First published on 5th February 2025


Abstract

Reduction of P4 with β-diketiminate MgI complexes, (BDI)MgMg(BDI), depends strongly on the bulk of the ligand. Whereas superbulky BDI ligands gave selective reduction to P42− in a butterfly conformation, reduction with a less bulky ligand gave various products among which P84− had a realgar-type structure. The selectivity of P4 reduction can also be controlled by metal choice. Reduction of P4 with CaI synthons of general type (BDI*)Ca–X–Ca(BDI*) in which BDI* is a superbulky ligand and X is a bridging dianion (C6H62− < p-xylene2− < N22−) led to reduction of P4 to the very common, stable Zintl anion P73−. Monitoring this process with 31P NMR shows that cyclo-P42− is an intermediate en route to P73−. Conversion rates increase with increasing reducing power: X = C6H62− < p-xylene2− < N22−. A complex with the weakly reducing DBA2− dianion led to selective P4 reduction to cyclo-P42− (DBA = 9,10-dimethyl-diboraanthracene). DBA inhibits P4-to-P7 conversion, most likely by capturing the electron needed for further P4 reduction by radical processes. Experimental investigations are supported by crystal structure determinations and a computational DFT study which also shows that the nature of metal–P4 bonding (covalent or ionic) determines the preference for formation of butterfly-shaped P42− or planar 6π-electron aromatic cyclo-P42−.


Introduction

White phosphorus (P4) is a commodity reagent for the production of industrially relevant P-containing products. While traditional bulk processes convert P4 with highly corrosive Cl2 to PCl3 for further functionalization with polar organo-metallic reagents, current research initiatives aim for catalytic protocols to directly convert P4 to organophosphorus compounds.1 In this light, the activation and chemical breakdown of P4 is an important research field. Being a highly strained molecule, P4 can be easily oxidized or reduced and shows diverse reactivity. Reacting either as an electrophile, nucleophile, or as an e-donor/acceptor, it could be seen as a chameleon in P-chemistry.2

Although P4 is inherently highly reactive, reaction pathways often remain unclear and selective conversions are difficult to achieve. Numerous groups have reported on P4 activation using the rich redox reactivity of the transition metals.3–6 Recent developments in low-valent p-block chemistry stimulated P4 activation with reagents that, due to small HOMO–LUMO gaps, show transition metal-like reactivity.7–9 Earlier highlights of this work include the insertion of (Me3Si)3CGa in three P–P bonds of P4 by Uhl and coworkers10 or the complete reduction of P4 by Cp*Al to give a P3− containing cluster (Cp*Al)6(P)4 (I, Scheme 1) by the Schnöckel group.11 In contrast, the bulkier β-diketiminate complex (BDI)Al reacted to give a complex of the P44− anion (II), formed by 4e-reduction and cleavage of two edges in the P4 tetrahedron; BDI is herein defined as HC[C(Me)-N(DIPP)]2 (DIPP = 2,6-diisopropylphenyl). Reaction with the softer reducing agent (BDI)Ga led only to 2e-reduction and cleavage of one P–P edge (III).12 The valence isoelectronic cyclic alkyl amino carbenes (CAACs) have also been shown to activate P4 (IV),13 whereas silylenes show either 4e-reduction giving P44−, like in II, or 2e-reduction resulting in P42−, like in III.14


image file: d4sc08502g-s1.tif
Scheme 1 Selected products of P4 activation by low-valent main group compounds.

Given the plethora of important breakthroughs in low-oxidation state s-block metal chemistry,15–17 it is remarkable that there is a complete lack of research on P4 reduction with these highly reactive early main group metal complexes. The reduction chemistry of one of the first MgI complexes, (BDI)MgMg(BDI),18 has been extensively investigated15 but we are unaware of reactivity studies with P4. However, this mild reducing agent has been reported to reduce the P5 ring in Cp*Fe(P5).19 The far majority of P4 activation studies with s-block metal reagents exploit their superb nucleophilicity. Classical examples include P–P bond cleavage in P4 by nucleophilic addition of RLi or Grignard reagents.20 An interesting case of nucleophilic activation of P4 is its reaction with nucleo-philic hydride reagents like [(BDI)Ca(μ2-H)]2 which after subsequent H2 release gave the reduction-like product [(BDI)Ca]3(P7), containing the Zintl P73− ion.19 In this example, the calcium hydride complex reacts as a synthon for the hitherto unknown CaI complex (BDI)CaCa(BDI). Such reductive reactivity of group 2 metal hydride complexes is well-established.21

We herein report a systematic study on the reduction of P4 with MgI complexes or CaI synthons, i.e. CaII complexes containing electron-rich ligands that can react like the corresponding CaI species.22−24 We demonstrate that selectivity is largely dependent on metal choice, ligand bulk or the nature of the electron-rich ligand delivering the electrons for P4 reduction.

Results and discussion

P4 activation with MgI complexes

The direct reduction of P4 with β-diketiminate MgI complexes of type (BDI)MgMg(BDI) has so far not been described in the literature. This could be because 1H and 31P{1H} NMR monitoring of an equimolar mixture of [(BDI)Mg]2 and P4 in C6D6 at room temperature showed a highly unselective conversion (Scheme 2 and Fig. S37/S38). Various side-reactions may originate from the poor solubility of the reactants and the heterogeneity of the reaction mixture. However, we found that gently heating the mixture for three days at 60 °C leads to further conversion and selective crystallization of a most insoluble reaction product [(BDI)Mg]4(P8) (1) in 10% yield (Fig. 1a). This minor product shows 31P NMR resonances at +68.3 and +145.0 ppm which were not observed in the crude product of the room temperature [(BDI)Mg]2/P4 conversion. This means that 1 was formed after thermal treatment. The P84− unit is isostructural and valence isoelectronic to α-P4S4, which is of the realgar-type,25 and flanked by four [(BDI)Mg]+ fragments at the corners with each Mg atom bound to two P atoms. Despite the high symmetry of the P84− anion, the complex shows no crystallographic symmetry. Although this structure is unprecedented in s-block metal chemistry, a few examples for P84− formation are known from transition metal,26–28 lanthanide,29 and gallium mediated P4 activation.30 Similar realgar-type polystibides Sb8 have been isolated as the corresponding [(BDI)Mg]4(Sb8) complexes with different BDI ligands.31 The Mg–P distances in 1 are in the narrow range of 2.599(1)–2.692(1) Å with P–Mg–P bite angles varying from 73.92(3)° to 74.79(3)°. In contrast to [(BDI)Mg]4(Sb)8 which exhibits an almost linear Mg–Sb–Mg arrangement, the Mg–P–Mg angles in 1 deviate slightly from linearity: 164.26(3)–168.05(3)°. Within the P84− anion in 1 there are two types of P–P bonds. The P–P bonds between three-coordinate P atoms, P2–P6 (2.2868(8) Å) and P4–P8 (2.2751(8) Å), are slightly elongated compared to the remaining P–P bonds ranging from 2.1999(7) to 2.2188(8) Å. The P–P–P angles range from 92.60(3)° to 102.67(3)°. These structural features are in agreement with those reported for the SmIII cluster (Cp*2Sm)4(P8).29
image file: d4sc08502g-s2.tif
Scheme 2 Activation of P4 by β-diketiminate MgI complexes.

image file: d4sc08502g-f1.tif
Fig. 1 (a) Crystal structure of [(BDI)Mg]4(P8) (1) in which iPr-groups and H atoms are omitted for clarity and a view of the Mg4P8 core. (b) Crystal structure of [(BDI*)Mg]2(P4) (2); the Et2CH-groups and H atoms are omitted for clarity. A crystallographic mirror plane runs through the atoms Mg1, Mg2, P2, P3.

In contrast, reaction of P4 with the more sterically hindered MgI complex [(BDI*)Mg]2,32 featuring a considerably elongated Mg–Mg interatomic distance,33 gave at room temperature overnight highly selective conversion; BDI* is defined as HC[C(Me)–N(DIPeP)]2 (DIPeP = 2,6-(Et2CH)-phenyl). In contrast to the unselective [(BDI)Mg]2/P4 conversion (Fig. S37/S38), the crude product of the [(BDI*)Mg]2/P4 conversion showed a 31P{1H} NMR spectrum with only two triplet resonances in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio (Fig. S41). Despite the high selectivity of this reaction, the very good solubility induced by the Et2CH-substituents allowed for isolation of crystalline [(BDI*)Mg]2(P4) (2) in only 34% yield.

In agreement with two triplet signals in 31P{1H} NMR, inspection of the crystal structure revealed a P42− dianion in a butterfly conformation which is bridging two (BDI*)Mg+ units in an unusual η22-fashion. Butterfly-shaped P42− anions usually bridge metals in η11-fashion.34 Recently, Aldridge and co-workers isolated an odd example of P42− bridging between Al and K in η21-fashion35 while Hill and co-workers reported bridging in η23-fashion.36 Complex 2 could also be considered to consist of a magnesate anion (BDI*)Mg(P4) with two polar Mg–P bonds to the two-coordinate P atoms (P2, P3) of 2.573(4)–2.694(4) Å, charge-balanced by a (BDI*)Mg+ cation that interacts with Mg–P interactions of circa 2.606(4) Å with both three-coordinate P-atoms (P1 and P1′). Due to disorder of the bridging P42− anion (Fig. S62) a more accurate discussion of the crystal structure is not possible.

P4 activation with CaI synthons

Although β-diketiminate stabilized CaI complexes are currently unknown, we reported a range of (BDI*)Ca–X–Ca(BDI*) complexes with various bridging X2− anions that react like a low-valent CaI complex (X = N2, benzene, p-xylene).22,24 Their reducing ability is directly related to the 2e-oxidation of the bridging anion, X2− → X + 2e, and therefore to the reduction potential of X.

Since the most reducing CaI synthon [(BDI*)Ca]2(N2) is unstable in aromatic solvents,22 the reaction with P4 was carried out in methylcyclohexane (Scheme 3). Reaction of the CaI synthon with a methylcyclohexane solution of P4 at room temperature led to immediate N2 evolution and a colour change from dark brown to dark orange. Analysis of the crude reaction mixture by 31P{1H} NMR spectroscopy revealed after 30 minutes a sharp low-field singlet at 458.4 ppm and a rather broad high-field singlet at −85.3 ppm in a ratio of 0.09[thin space (1/6-em)]:[thin space (1/6-em)]0.91, respectively (Fig. S42). Continued stirring for another 4 hours showed the selective formation of one species corresponding to the broad high-field signal at −85.3 ppm (Fig. S43). Layering a saturated methylcyclohexane/pentane solution with a few drops of tetrahydropyran (THP) resulted in the isolation of yellow crystals of composition [(BDI*)Ca(THP)]3(P7) (3) in a 28% yield.


image file: d4sc08502g-s3.tif
Scheme 3 Activation of P4 by CaI synthons.

The crystal structure of 3 (Fig. 2a) revealed that P4 had been reduced to the polyphosphide P73− Zintl ion, encapsulated by three [(BDI*)Ca(THP)]+ fragments. A similar product, [(BDI)Ca]3(P7), has been obtained by Roesky and co-workers by reduction of P4 with [(BDI)Ca(μ2-H)]2.19 However, in contrast to this previous report which describes a major side-product with a 31P-resonance at −241.3 ppm, the reduction of P4 with the CaI synthon is highly selective.


image file: d4sc08502g-f2.tif
Fig. 2 (a) Crystal structure of [(BDI*)Ca(THP)]3(P7) (3) and a view of the Ca3P7 core. (b) Centrosymmetric crystal structure of [(BDI*)Ca]2(DBA) (4). (c) Centrosymmetric crystal structure of [(BDI*)Ca(OEt2)]2(cyclo-P4) (5). In all figures the H atoms and Et2CH-groups have been omitted for clarity.

Although the structure of 3 is close to being C3-symmetric, the complex does not show crystallographic symmetry. The Ca metal centres have shortest interactions to the three two-coordinate P atoms (P4–P6) in P73− which formally carry a negative charge (Ca–P: 3.0557(7)–3.1163(6) Å). These bonds are considerably longer than the corresponding Ca–P bonds in [(BDI)Ca]3(P7) which vary from 2.8667(9) to 2.9346(9) Å.19 The long Ca–P contacts in 3 are mainly due to the bulkier BDI* ligand and the additional THP coordination. While in 3 there are no Ca–P contacts to the apical P atom P7, the formally neutral three-coordinate P atoms in the P3-triangle (P1–P3) show long contacts (Ca–P: 3.3253(7)–3.4611(5) Å). The P–P bond lengths in the P7-cage (2.1601(6)–2.2610(4) Å) are in the range of previous reported Zintl ions.37 Reports of group 2 metal based Zintl ions are particularly rare and especially their selective formation from P4 remains difficult. Hill and co-workers isolated a Mg flanked P7 Zintl cluster that is structurally very similar to [(BDI)Ca]3(P7). However, this product could only be obtained in poor yields as a minor side product by fractional crystallization of the raw product.38

Whilst the 31P{1H} NMR spectrum of [(BDI*)Ca(THP)]3(P7) at room temperature showed one very broad singlet at −85.28 ppm due to fast exchange between the three different P-positions in the P73− anion, at −90 °C this resonance was split into seven broad but distinct signals (Fig. S22), partially with visible magnetic coupling (ppm: −147.6, −130.4, −111.6, −75.7, −69.2, −45.1, and −38.2). This behaviour differs from other reports on the dynamics of P73− which generally describe splitting in only three signals in a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]3[thin space (1/6-em)]:[thin space (1/6-em)]3 upon cooling.19,37 These can be assigned to the apical position, the P3 basal triangle and the three connecting P atoms. Our observed splitting into seven signals at low temperature can only be rationalized by a loss of trigonal symmetry by different coordination geometries at the Ca centres, making each P atom magnetically inequivalent.

The central 2e-donor X in (BDI*)Ca–X–Ca(BDI*) influences the synthon's stability and reactivity.22−24 In order to evaluate the effect of the central X2− anion on P4 reduction, the most reducing CaI synthon, [(BDI*)Ca]2(N2) was converted to the corresponding arene complexes by reaction with benzene or p-xylene which led to N2 release. These binuclear arene complexes were further reacted with equimolar quantities of P4 in cyclohexane-d12 and conversion was monitored with 31P{1H} NMR. Similarly to P4 reduction with N22−, two signals were observed: a sharp low-field resonance at 458.4 ppm and a rather broad high-field singlet at −85.3 ppm, corresponding to the P73− Zintl anion. Over time, the sharp low-field resonance at 458.4 ppm disappeared and clean formation of the P7 complex was observed. This suggests that the species with a 31P resonance at 458.4 ppm is an intermediate on the way to formation of the P7 complex. The only difference is the rate at which this happens which seems to be related to the reducing power of the central X2− anion: N22− > p-xylene2− > benzene2−. While in the case of the N2 complex, the intermediate species was nearly fully converted to the P7 complex after an hour, the less reducing benzene complex required stirring overnight (Fig. S45). As we were not able to isolate the intermediate, we sought a weaker CaI synthon to slow the reaction further.

Reaction of P4 with (BDI*)Ca–(anthracene)–Ca(BDI*)23 gave a myriad of products of which one could be recognized as the P7 complex (3) by 31P NMR analysis (Fig. S48). This prompted us to look further for a suitable bridging ligand. Due to aromaticity in its dianionic state, the boron-doped 9,10-dimethyl-diboraanthracene (DBA) dianion is less reducing and has markedly different electronic properties.39–41 Recently, first lanthanide triple-decker complexes featuring DBA2− ligands were reported.42 Reduction of DBA with the CaI synthon [(BDI*)Ca](N2) in cyclohexane resulted in gas evolution and immediate precipitation of a microcrystalline orange solid (Scheme 3). 1H NMR showed selective formation of the target complex [(BDI*)Ca]2(DBA) (4) which could be isolated in 88% yield (Fig. S49). Unlike (BDI*)Ca–(C6H6)–Ca(BDI*), which in benzene shows reductive benzene coupling (C6H62− + C6H6 → biphenyl2−),43 [(BDI*)Ca]2(DBA) is even at 60 °C remarkably stable in aromatic solvents. Its 11B NMR spectrum did not show a clear signal, even when quartz NMR tubes were used. Bright orange crystals suitable for single crystal X-ray diffraction were obtained by recrystallization from a cyclohexane/n-pentane mixture at room temperature.

The complex crystallized in the P[1 with combining macron] space group with two independent [(BDI*)Ca]2(DBA) (4) inverse sandwich complexes in the asymmetric unit (Fig. 2b). The structure shows a planar (μ266-DBA)2− ligand of which the central B2C4-ring is sandwiched between the two (BDI*)Ca+ fragments with Ca–ringcentre distances ranging from 2.3552(4) to 2.3786(4) Å. Preference for metal coordination to the central B2C4-ring was also observed in alkali metal complexes: M(THF)n–(DBA)–M(THF)n complexes (M = Li, Na, K).40,44 Like in these alkali metal inverse sandwiches, the planar DBA2− anion in 4 is isoelectronic to anthracene and shows in its central ring C–C bond distances of 1.466(3)–1.468(2) Å and B–C bond distances of 1.530(3)–1.541(3) Å, specifying the extended aromatic nature of this dianion. Computational investigation of 4 shows that Ca ligand bonding is highly ionic. NPA charges: (BDI*) −0.88, Ca +1.79, DBA −1.87 (Fig. S67).

A C6D6 solution of [(BDI*)Ca]2(DBA) (4) and one equivalent of P4 was stirred overnight at room temperature resulting in a colour change from orange to bright yellow. The 31P{1H} NMR spectrum showed a sharp singlet at 453.9 ppm indicative for exclusive formation of the hitherto unidentified intermediate (Fig. S51) whereas the high field signal at −85 ppm for P73− formation is missing. In addition, 1H, 11B, and 13C NMR confirmed the release of neutral DBA (Fig. S52–S54). The sharp low-field singlet of this intermediate could be assigned to the four chemically equivalent phosphorus atoms of the cyclo-P4 dianion, in the form of an inverse sandwich complex. Comparing to few reports in literature, the 31P NMR chemical shift of the cyclo-P42− ring is sensitive to its environment: [(DIPPForm)2Sm]2(cyclo-P4) with δ = +453 ppm45 (DIPPForm = HC(N-DIPP)2), Cs2P4·2NH3 with δ = +348 ppm,46 [(NON)Sm(THF)2]2(cyclo-P4) with δ = +480 ppm (NON = 4,5-bis(2,6-diisopropylphenyl-anilido)-2,7-di-tert-butyl-9,9-dimethyl-xanthene),47 [(NON)Yb(THF)2]2(cyclo-P4) with δ = +382 ppm,47 and a U complex with η22-bridging cyclo-P42− with δ = +718 ppm.48 Despite this considerable range in chemical shifts, all values are considerably downfield shifted, substantiating the 6π-electron aromatic character of this dianion.

Crystallization from a saturated pentane solution, layered with drops of Et2O, allowed for isolation and structural characterization of [(BDI*)Ca(OEt2)]2(cyclo-P4) (5) by X-ray diffraction. The crystal structure of 5 confirms the formation of an inverse sandwich complex with a bridging cyclo-P4 dianion (Fig. 2c). The 6π-electron aromatic cyclo-P42− ring bridges in η44-fashion between two cationic (BDI*)Ca+ fragments with Ca–P4(centroid) bond lengths of 2.6460(5) Å and Ca–P bond lengths between 3.0446(9) and 3.0682(8) Å. The P4 ring is disordered over two positions which are rotated in respect to each other around the Ca⋯Ca′ axis by circa 45°. The P–P bond lengths range from 2.156(1) Å to 2.158(1) Å and internal P–P–P angles are close to 90° (P2–P1–P2′ 89.40(4)°, P1′–P2–P1 90.60(4)°). Hence, the aromatic P42− ring is almost perfectly square planar. The P–P bond lengths of the phosphorus ring are in the range of reported examples and seems independent of the metal atoms which sandwiches this ring (Sm: 2.144(1)–2.162(1) Å,45 U: 2.149(2)–2.152(2) Å,48 Cs: 2.146(1)–2.148(1) Å).46 It is of interest to note that the geometry of P42− depends strongly on the metals that sandwich this dianion. The ring structure of 6π-electron aromatic cyclo-P42− in the Ca inverse sandwich 5 contrasts strongly with the butterfly structure of P42− in the Mg inverse sandwich 2. A detailed discussion on these differences follows below (vide infra).

It is noteworthy, that the reduction of P4 with [(BDI*)Ca]2(DBA) (4) exclusively led to formation of the aromatic (cyclo-P4)2− dianion. Even with a large excess of P4 and using forcing reaction conditions (60 °C), only the P4 complex 5 was formed. This represents the first example of a quantitative reduction of P4 to (cyclo-P4)2− mediated by s-block metals. Treatment of P4 with Cs in THF followed by solvation in liquid ammonia led to the formation of Cs3P7·3NH3 as major product and the desired cyclotetraphosphide Cs2P4·2NH3 was only a by-product.46

In order to evaluate its role as an intermediate in polyphosphide formation, crystalline [(BDI*)Ca(OEt2)]2(cyclo-P4) (5) and one equivalent of P4 were suspended in cyclohexane-d12 (Scheme 3). Monitoring the reaction with 31P{1H} NMR showed after 2 hours at room temperature slow conversion of the (cyclo-P4)2− complex into the (P7)3− complex (P4[thin space (1/6-em)]:[thin space (1/6-em)]P7 = 0.63[thin space (1/6-em)]:[thin space (1/6-em)]0.37). Complete, selective conversion to the P7 product was achieved after two days at room temperature (Fig. S55).

Two observations need further attention. (1) The DBA complex 4 cannot be converted to a P7 complex and reacts with excess white phosphorous only to the P4 product. (2) Once crystallized in presence of ether and isolated, 5 selectively reacts with P4 to the P7 complex. Although at first sight contradicting, these combined observations can only lead to one conclusion. The neutral DBA that is released in reaction of 4 with P4 must be an inhibitor for the P4-to-P7 conversion.

Indeed, whereas the reaction of crystalline [(BDI*)Ca(OEt2)]2(cyclo-P4) (5) with P4 in cyclohexane-d12 showed slow but selective conversion to the P7 complex, addition of DBA to 5 inhibited P4-to-P7 conversion. With catalytic quantities of DBA as low as 5 mol% no conversion was observed, even after one day at room temperature. However, after 30 hours at 60 °C a small amount of the P7 complex was observed (Fig. S58). The mechanism of this inhibitor effect are still unclear but leave room for speculation (vide infra).

P42−: butterfly or aromatic ring?

The geometry of P42− depends strongly on the metals that sandwich this dianion. Captured between two (BDI*)Mg+ cations it takes the form of a butterfly with η22-bridging (2) but between (BDI*)Ca+ a η44-bridging 6π-electron aromatic cyclo-P42− dianion (5) is favoured. These intriguing differences in geometry and coordination modes may be understood by Density Functional Theory (DFT) calculations.

The structures of 2 and ether-free 5 were optimized at the PBE0/def2-SVP level of theory. The calculated structure of 2 fits reasonably well with that from the crystal structure (Fig. S66), indicating a sufficient level of theory. The aromaticity of 5 has been investigated before, and differs from that in classical aromatic hydrocarbons like in benzene.49 It can be noticed that the P–P bonds in cyclo-P42− in 5 (calculated: 2.173–2.184 Å) are not much different from the single P–P bonds in P4 (calculated: 2.201 Å). Electron Localization Function (ELF) analyses showed that the P–P bonds in cyclo-P42− have P–P single bond character and show no sign of the high electron delocalization as found in benzene.49 In contrast to the high population in C–C bonds of benzene, ELF analysis does not show a high population in the P–P bonding orbitals. Instead, a high population was found in the P lone-pairs in cyclo-P42−. Hence, the concept of “lone-pair aromaticity” was defined. In contrast to benzene, for which a circle within the C6-ring depicts aromaticity, for cyclo-P42− a circle around the P4-ring has been proposed.49

The question remains why the P42− dianion can have two different appearances, butterfly or cyclic structure, depending on the metal that sandwich this entity. The answer may lie in the character of metal–P4 bonding. Based on electronegativity differences (Mg: 1.31, Ca: 1.00, P: 2.19),50 Mg–P4 bonding in 2 should be slightly more covalent than Ca–P4 bonding in 5. The rather low Wiberg Bond Indices (WBI's) for Mg–P4 bonding (0.11/0.17) and Ca–P4 bonding (0.05/0.06) show that both bonds have ionic character but the Mg–P4 bonds are somewhat more covalent (Fig. 3a). Natural Population Analysis (NPA) indeed shows a significantly higher negative charge on the cyclo-P4 unit in Ca complex 5 (−1.74) than that on butterfly P4 unit in Mg complex 2 (−1.56); see Fig. 3a. This corresponds to higher positive charges on the Ca cations (+1.74) compared to the Mg cations (average: +1.67). The butterfly form of P42− shows two different P atoms. The two-coordinate P atoms that carry most of the negative charge (−0.55) shows strong bonding to Mg and feature WBI's of 0.17. The three-coordinate P atoms with charges of −0.23 have weaker bonds to Mg (WBI: 0.11). This confirms the view that 2 can be seen as a magnesate anion (BDI*)Mg(P4) with a total NPA charge of −0.82 that interacts with a (BDI*)Mg+ cation carrying a charge of +0.82. The WBI's for the P–P bonds (0.94/0.98) are close to those for a covalent single bond.


image file: d4sc08502g-f3.tif
Fig. 3 Computational studies at the PBE0-D3BJ(PCM = cyclohexane)/def2-TZVP level of theory. (a) NPA charges and WBI's for [(BDI*)Mg]2(P4) and [(BDI*)Ca]2(cyclo-P4). (b) Energy profile for the reaction of [(BDI*)Ca]2(cyclo-P4) with DBA.

In comparison, the NPA charges on the P atoms in cyclo-P4 are more similar (−0.38/−0.49) and intermediate to those in butterfly P42− (−0.23/−0.55). Charge differences are likely dictated by their different environments. Atoms-In-Molecules (AIM) analyses show that the P42− dianions in 2 and 5 are also involved in weak P⋯H–C bonding with organic fragments of the BDI* ligands (Fig. S72 and S74). In agreement with more ionic character of the Ca complex, the WBI's for Ca–P4 bonding (0.05/0.06) are smaller than for Mg–P4 bonding (0.11/0.17). However, the WBI's for the P–P bonds in cyclo-P4 (1.23/1.25) are slightly larger than that expected for a single bond. This is in agreement with some extent of aromaticity (vide supra).

The differences observed in bonding of the P42− dianion in Mg and Ca complexes are comparable to differences observed in bonding of the C6H62− dianion (V and VI, Scheme 4). Bonding in the Mg complex (BDI*)Mg–(C6H6)–Mg(BDI*) is more covalent than that in the corresponding Ca complex. The C6H62− in the Mg complex shows the typical boat form with strong localized Mg–C bonding to bow and stern and much weaker Mg–C interactions to the C[double bond, length as m-dash]C bonds.33 Similarly as for 2, the complex can be seen as a magnesate anion (BDI*)Mg(C6H6) interacting with a (BDI*)Mg+ cation through unusual Mg–alkene coordination for which we recently found ample proof.51–53 In contrast, (BDI*)Ca–(C6H6)–Ca(BDI*) shows a nearly flat C6H62− dianion and η66-bridging between the Ca2+ ions.22 These parallels between C6H62− and P42− bonding in Mg or Ca sandwich complexes find their origins in the more covalent nature of the Mg–ligand bond but could also be related to the considerably larger size of the Ca2+ cation compared to the Mg2+ cation. This explanation is in agreement with the occurrence of P42− butterfly structures in Al, Ga, Si or Ni complexes which are even more covalent in character than the Mg complex.12,14,34–36


image file: d4sc08502g-s4.tif
Scheme 4 Formulas V–VIII.

How does diboraanthracene (DBA) inhibit P4 to P7 conversion?

Whilst [(BDI*)Ca(OEt2)]2(cyclo-P4) (5) reacts selectively with P4 to the P7 complex, addition of DBA inhibited this conversion. Initially, we sought an explanation in possible interaction of DBA with the P4 reactant. It is known that the electrophilic diborane DBA can interact with bidentate electron-rich ligands like 1,2-diazines (e.g. in VII)54 and even assists in N2 fixation with Cp*2Sm (VIII) (Cp* = 1,2,3,4,5-penta-methylcyclopentadienyl).55 However, 1H and 31P NMR spectra of a mixture of DBA and P4 did not show any changes in the chemical shifts when compared to the pure species. This excludes significant DBA⋯P4 interaction. Attempts to optimize the structure of potential DBA⋯P4 complexes by DFT calculation also only led to separation of these molecules.

Alternatively, DBA can interact with the P4 complex [(BDI*)Ca]2(cyclo-P4), thus inhibiting further reactivity with P4. DFT calculations indeed show that this complex is able to interact with DBA by formation of a P–B bond (Fig. 3b; for structures see Fig. S78). Although the activation free energy for this process is only ΔG298 = 12.5 kcal mol−1, the adduct is ΔG298 = 7.8 kcal mol−1 higher in energy than the unbound molecules. This implies a fast equilibrium that lies mainly on the side of free P4 and [(BDI*)Ca]2(cyclo-P4). This means that, although complexation of DBA with cyclo-P42− is possible, this cannot be a reason to inhibit further reactivity towards the formation of P7 complex. Especially, when one considers that also small, catalytic quantities of DBA already work as inhibitor.

Reaction mechanisms that give rise to higher nuclearity clusters are in general poorly understood.37 However, it seems reasonable to assume that reduction of P4 with s-block metal reducing agents starts with electron-transfer. As we could show that the P42− dianion is a likely intermediate en route to P73−, the first step of this transformation could be a single electron transfer (SET) process. As DBA is a molecule with a low-lying LUMO and can be easily reduced, its mode of inhibiting the P4-to-P7 conversion may simply be a reversible electron-capture process. The high affinity of DBA for electrons inhibits SET to P4. Using DFT, we found that reduction of DBA is indeed much more facile than reduction of P4. The reaction, DBA + P4 → P4 + DBA, was calculated to be highly endothermic (in cyclohexane: ΔH = +23.5 kcal mol−1). This is also in agreement with the observation that catalytic quantities of DBA function as a stabilizer for the P42− dianion and would explain the highly selective formation of the P4 complex 5.

Conclusion

We achieved first P4 reductions with low-valent β-diketiminate MgI complexes and found the selectivity to be greatly dependent on the bulk of the ligand. Using a superbulky BDI ligand with DIPeP substituents (BDI*) led to highly selective formation of a butterfly-shaped P42− dianion bridging two (BDI*)Mg+ fragments in a unique η22-fashion. The same complex with Ca instead of Mg features a cyclo-P42− dianion bridging (BDI*)Ca+ fragments in an η44 mode. This difference finds its origin in the more covalent nature of the Mg–(P4) bond but could also be related to the considerably larger size of the Ca2+ cation compared to the Mg2+ cation, favouring a delocalized cyclo-P42− dianion.

Reaction of low-valent CaI synthons selectively gave products with the P73− Zintl anion: [(BDI*)Ca]3(P7). Monitoring these reactions by NMR shows first unambiguous proof that P73− formation proceeds through a cyclo-P42− intermediate. The kinetics of the P4cyclo-P42− → P73− conversion depends strongly on the reducing power of the CaI synthon: [(BDI*)Ca]2(X) in which X is dianionic N22− or arene2−. Conversion is faster along the row X = benzene2− < p-xylene2− < N22−.

In case of CaI synthons, using 9,10-dimethyl-diboraanthracene (DBA) as a bridging dianion led to exclusive formation of the cyclo-P42− product. Even with excess P4 and forcing reaction conditions no further P4 → P7 conversion was observed in the time frame conducted. As addition of trace quantities of DBA already inhibited and considerably retarded further reactivity of the cyclo-P42− dianion, it is proposed that DBA prevents radical reactivity by functioning as a reversible electron trap.

These first investigations on P4 reduction with low-valent Ae metal complexes show that selectivities depend on the bulk of the BDI ligand, the metal and the presence of inhibitors for radical reactivity. We continue our research with investigations on P4 reduction with heavier low-valent Ae metal synthons.

Data availability

Crystallographic data has been deposited with the Cambridge Structural Database.

Author contributions

S. Thum: conceptualization, investigation, validation, formal analysis, writing – original draft, visualization. O. P. E. Townrow: investigation, validation, formal Analysis. J. Langer: formal analysis, validation. Sjoerd Harder: conceptualization, writing – original draft – review and editing, visualization, validation, supervision, project administration.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We acknowledge Mrs A. Roth (University of Erlangen-Nürnberg) for CHN analyses and J. Schmidt and Dr C. Färber (University of Erlangen-Nürnberg) for assistance with the NMR analyses. O. P. E. T. thanks the Alexander von Humboldt Foundation for a postdoctoral fellowship.

References

  1. U. Lennert, P. B. Arockiam, V. Streitferdt, D. J. Scott, C. Rödl, R. M. Gschwind and R. Wolf, Nat. Catal., 2019, 2, 1101–1106 CrossRef CAS PubMed.
  2. D. J. Scott, Angew. Chem., Int. Ed., 2022, 61, e202205019 Search PubMed.
  3. C. M. Hoidn, D. J. Scott and R. Wolf, Chem.–Eur. J., 2021, 27, 1886–1902 Search PubMed.
  4. B. M. Cossairt, N. A. Piro and C. C. Cummins, Chem. Rev., 2010, 110, 4164–4177 CrossRef CAS PubMed.
  5. M. Caporali, L. Gonsalvi, A. Rossin and M. Peruzzini, Chem. Rev., 2010, 110, 4178–4235 CrossRef CAS PubMed.
  6. L. Giusti, V. R. Landaeta, M. Vanni, J. A. Kelly, R. Wolf and M. Caporali, Coord. Chem. Rev., 2021, 441, 213927 CrossRef CAS.
  7. M. Scheer, G. Balázs and A. Seitz, Chem. Rev., 2010, 110, 4236–4256 CrossRef CAS PubMed.
  8. S. Khan, S. S. Sen and H. W. Roesky, Chem. Commun., 2012, 48, 2169 RSC.
  9. N. A. Giffin and J. D. Masuda, Coord. Chem. Rev., 2011, 255, 1342–1359 CrossRef CAS.
  10. W. Uhl and M. Benter, Chem. Commun., 1999, 771–772 RSC.
  11. C. Dohmeier, H. Schnöckel, C. Robl, U. Schneider and R. Ahlrichs, Angew. Chem., Int. Ed., 1994, 33, 199–200 CrossRef.
  12. G. Prabusankar, A. Doddi, C. Gemel, M. Winter and R. A. Fischer, Inorg. Chem., 2010, 49, 7976–7980 CrossRef CAS PubMed.
  13. J. D. Masuda, W. W. Schoeller, B. Donnadieu and G. Bertrand, Angew. Chem., Int. Ed., 2007, 46, 7052–7055 Search PubMed.
  14. Y. Xiong, S. Yao, M. Brym and M. Driess, Angew. Chem., Int. Ed., 2007, 46, 4511–4513 CrossRef CAS PubMed.
  15. C. Jones, Nat. Rev. Chem, 2017, 1, 0059 CrossRef CAS.
  16. B. Rösch and S. Harder, Chem. Commun., 2021, 57, 9354–9365 RSC.
  17. L. A. Freeman, J. E. Walley and R. J. Gilliard, Nat. Synth., 2022, 1, 439–448 CrossRef CAS.
  18. S. P. Green, C. Jones and A. Stasch, Science, 2007, 318, 1754–1757 Search PubMed.
  19. R. Yadav, M. Weber, A. K. Singh, L. Münzfeld, J. Gramüller, R. M. Gschwind, M. Scheer and P. W. Roesky, Chem.–Eur. J., 2021, 27, 14128–14137 Search PubMed.
  20. M. M. Rauhut and A. M. Semsel, J. Org. Chem., 1963, 28, 471–473 Search PubMed.
  21. A. S. S. Wilson, C. Dinoi, M. S. Hill, M. F. Mahon, L. Maron and E. Richards, Angew. Chem., Int. Ed., 2020, 59, 1232–1237 CrossRef CAS PubMed.
  22. B. Rösch, T. X. Gentner, J. Langer, C. Färber, J. Eyselein, L. Zhao, C. Ding, G. Frenking and S. Harder, Science, 2021, 371, 1125–1128 CrossRef PubMed.
  23. J. Mai, B. Rösch, N. Patel, J. Langer and S. Harder, Chem. Sci., 2023, 14, 4724–4734 RSC.
  24. J. Mai, B. Rösch, J. Langer, S. Grams, M. Morasch and S. Harder, Eur. J. Inorg. Chem., 2023, 26, e202300421 CrossRef CAS.
  25. A. M. Griffin, P. C. Minshall and G. M. Sheldrick, J. Chem. Soc., Chem. Commun., 1976, 809–810 RSC.
  26. W. Huang and P. L. Diaconescu, Chem. Commun., 2012, 48, 2216–2218 RSC.
  27. M. E. Barr, B. R. Adams, R. R. Weller and L. F. Dahl, J. Am. Chem. Soc., 1991, 113, 3052–3060 Search PubMed.
  28. F. Spitzer, C. Graßl, G. Balázs, E. M. Zolnhofer, K. Meyer and M. Scheer, Angew. Chem., Int. Ed., 2016, 55, 4340–4344 Search PubMed.
  29. S. N. Konchenko, N. A. Pushkarevsky, M. T. Gamer, R. Köppe, H. Schnöckel and P. W. Roesky, J. Am. Chem. Soc., 2009, 131, 5740–5741 CrossRef CAS PubMed.
  30. F. Hennersdorf, J. Frötschel and J. J. Weigand, J. Am. Chem. Soc., 2017, 139, 14592–14604 CrossRef CAS PubMed.
  31. C. Ganesamoorthy, C. Wölper, A. S. Nizovtsev and S. Schulz, Angew. Chem., Int. Ed., 2016, 55, 4204–4209 CrossRef CAS PubMed.
  32. T. X. Gentner, B. Rösch, K. Thum, J. Langer, G. Ballmann, J. Pahl, W. A. Donaubauer, F. Hampel and S. Harder, Organometallics, 2019, 38, 2485–2493 Search PubMed.
  33. T. X. Gentner, B. Rösch, G. Ballmann, J. Langer, H. Elsen and S. Harder, Angew. Chem., Int. Ed., 2019, 58, 607–611 CrossRef CAS PubMed.
  34. S. Pelties, A. W. Ehlers and R. Wolf, Chem. Commun., 2016, 52, 6601–6604 RSC.
  35. M. M. D. Roy, A. Heilmann, M. A. Ellwanger and S. Aldridge, Angew. Chem., Int. Ed., 2021, 60, 26550–26554 CrossRef CAS PubMed.
  36. M. S. Hill, M. F. Mahon, C. L. McMullin, S. E. Neale, K. G. Pearce and R. J. Schwamm, Z. Anorg. Allg. Chem., 2022, 648, e202200224 Search PubMed.
  37. R. S. P. Turbervill and J. M. Goicoechea, Chem. Rev., 2014, 114, 10807–10828 Search PubMed.
  38. M. Arrowsmith, M. S. Hill, A. L. Johnson, G. Kociok-Köhn and M. F. Mahon, Angew. Chem., Int. Ed., 2015, 54, 7882–7885 CrossRef CAS PubMed.
  39. M. Dietz, M. Arrowsmith, K. Drepper, A. Gärtner, I. Krummenacher, R. Bertermann, M. Finze and H. Braunschweig, J. Am. Chem. Soc., 2023, 145, 15001–15015 Search PubMed.
  40. E. Von Grotthuss, S. E. Prey, M. Bolte, H. W. Lerner and M. Wagner, J. Am. Chem. Soc., 2019, 141, 6082–6091 CrossRef CAS PubMed.
  41. S. E. Prey and M. Wagner, Adv. Synth. Catal., 2021, 363, 2290–2309 CrossRef CAS.
  42. C. Uhlmann, L. Münzfeld, A. Hauser, T. T. Ruan, S. Kumar Kuppusamy, C. Jin, M. Ruben, K. Fink, E. Moreno-Pineda and P. W. Roesky, Angew. Chem., Int. Ed., 2024, 63, e202401372 CrossRef CAS PubMed.
  43. J. Mai, M. Morasch, D. Jędrzkiewicz, J. Langer, B. Rösch and S. Harder, Angew. Chem., Int. Ed., 2023, 62, e202212463 CrossRef CAS PubMed.
  44. E. von Grotthuss, S. E. Prey, M. Bolte, H. W. Lerner and M. Wagner, Angew. Chem., Int. Ed., 2018, 57, 16491–16495 CrossRef CAS PubMed.
  45. C. Schoo, S. Bestgen, R. Köppe, S. N. Konchenko and P. W. Roesky, Chem. Commun., 2018, 54, 4770–4773 RSC.
  46. F. Kraus, J. C. Aschenbrenner and N. Korber, Angew. Chem., Int. Ed., 2003, 42, 4030–4033 CrossRef CAS PubMed.
  47. A. Hauser, L. Münzfeld, S. Schlittenhardt, R. Köppe, C. Uhlmann, U.-C. Rauska, M. Ruben and P. W. Roesky, Chem. Sci., 2023, 14, 2149–2158 RSC.
  48. A. S. P. Frey, F. G. N. Cloke, P. B. Hitchcock and J. C. Green, New J. Chem., 2011, 35, 2022–2026 RSC.
  49. F. Kraus and N. Korber, Chem.–Eur. J., 2005, 11, 5945–5959 CrossRef CAS PubMed.
  50. L. R. Murphy, T. L. Meek, A. Louis Allred and L. C. Alien, J. Phys. Chem. A, 2000, 104, 5867–5871 CrossRef CAS.
  51. J. Martin, J. Langer, M. Wiesinger, H. Elsen and S. Harder, Eur. J. Inorg. Chem., 2020, 2582–2595 CrossRef CAS.
  52. K. Thum, A. Friedrich, J. Pahl, H. Elsen, J. Langer and S. Harder, Chem.–Eur. J., 2021, 27, 2513–2522 CrossRef CAS PubMed.
  53. K. Thum, J. Pahl, J. Eyselein, H. Elsen, J. Langer and S. Harder, Chem. Commun., 2021, 57, 5278–5281 RSC.
  54. S. N. Kessler and H. A. Wegner, Org. Lett., 2010, 12, 4062–4065 CrossRef CAS PubMed.
  55. S. Xu, L. A. Essex, J. Q. Nguyen, P. Farias, J. W. Ziller, W. H. Harman and W. J. Evans, Dalton Trans., 2021, 50, 15000–15002 RSC.

Footnote

Electronic supplementary information (ESI) available: Experimental details, 1H and 13C NMR spectra, crystallographic details including ORTEP presentations, details for the DFT calculations including XYZ-files. CCDC 2407627–2407631. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4sc08502g

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.