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

Trends in benzene inverse sandwich complexes of the alkaline-earth metals Mg, Ca, Sr and Ba

Dawid Jędrzkiewicz *a, Michael Morasch a, Oliver P. E. Townrow ab, Bastian Rösch a, Jens Langer a, Zachary Mathe c and Sjoerd Harder *a
aInorganic and Organometallic Chemistry, Friedrich-Alexander-Universität Erlangen-Nürnberg, Egerlandstraße 1, 91058 Erlangen, Germany. E-mail: dawid.jedrzkiewicz@gmail.com; sjoerd.harder@fau.de
bInstitute of Nanotechnology, Karlsruher Institut für Technologie, Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany
cDepartment of Inorganic Spectroscopy, Max Planck Institute for Chemical Energy Conversion, Stiftstraße 34-36, 45470 Mülheim an der Ruhr, Germany

Received 18th July 2025 , Accepted 11th August 2025

First published on 29th August 2025


Abstract

Mechanochemical reduction of β-diketiminate (BDI) barium iodide precursors with K/KI resulted in the first barium inverse sandwich complexes containing the benzene dianion in yields of up to 54%. This most challenging isolation of highly reactive (BDI)Ba–(C6H6)–Ba(BDI) complexes, completes the family of heavier benzene inverse sandwich complexes and allows for a comparison of trends in the series from Mg, Ca, Sr to Ba. Syntheses, stabilities, structures, electronic states and reactivities of the full range are compared. Crucial for isolation of the Ba inverse sandwich complexes are the tBu-substituents in the ligand backbone which push the bulky aryl rings towards the large Ba metal cations. These secondary Ba⋯(π-Ar) interactions result in an unexpected high stability. Another trend is found for the ring puckering in the bridging benzene2− dianion which steadily increases from Ba to Mg. DFT calculations show the general ionic character of (BDI)Ae–(C6H6)–Ae(BDI) complexes (Ae = Mg, Ca, Sr, Ba) and reveal only small energy differences between closed-shell singlet or open-shell triplet states. The most reactive (BDI)Ba–(C6H6)–Ba(BDI) complexes could be considered the first BaI synthons. They reduce a range of polyaromatic hydrocarbons, H2 or even convert (BDI)MgI precursors into well-known (BDI)Mg–Mg(BDI) complexes. Reactions with heavier (BDI)AeI (Ae = Ca, Sr) gave (BDI)Ae–(C6H6)–Ae(BDI) and (BDI)BaI.


Introduction

Recent developments in low oxidation state alkaline earth metal (Ae) chemistry have expanded beyond the well-established dimeric magnesium(I) complexes that find widespread applications as precise reducing agents.1–4 The most significant synthetic milestones include isolation of the first complex with a BeI–BeI bond (I, Scheme 1),5 unique Mg0 complexes,6 and a range of derivatives featuring heterobimetallic Mg–Be, Mg–Ca, Mg–Sr, and Mg–Ba bonding (II–III, Scheme 1).7,8 Based on the decreasing electronegativity along the row Be > Mg > Ca > Sr > Ba, the formal oxidation states in II are MgII–Be0 whereas those in III are Mg0–AeII. Examples of heavier, homometallic AeI–AeI complexes are hitherto unknown. Computational analysis predicts that the Ca–Ca bond is much weaker than existing Be–Be and Mg–Mg bonds, making it too reactive to be isolable.9–12 This trend is expected to continue down the group.13,16 Nonetheless, the reactivity of heavier AeI complexes can be studied by means of AeI synthons, i.e. formal AeII complexes showing the reactivity of electron-rich AeI reagents. Whereas complexes with Be–Be or Mg–Mg bonds have a non-nuclear attractor (NNA), i.e. a local maximum of electron density, at the centre of the Be–Be and Mg–Mg axes, AeI synthons typically feature a molecular 2e reservoir that bridges the two AeII centres. Examples of such electron reservoirs include the dinitrogen dianion N22− (IV, Scheme 1)9,14–16 and the benzene dianion C6H62− (V, Scheme 1).9,17–20 Both of these dianions release their additional electrons easily, which makes them very strong reductants but they can also act as Brønsted bases or nucleophiles.6,18,19
image file: d5sc05373k-s1.tif
Scheme 1 Recently reported compounds containing Ae–Ae or Mg–Ae bonds (I–III) and examples of AeI synthons with N22− (IV) or C6H62− (V) electron reservoirs.

The strongly reducing character of these AeI synthons is determined by the very negative reduction potentials of N2 or C6H6. However, we found that the metals also have a strong influence. Their reactivities sharply increase with the size of the Ae metal. Longer and more ionic Ae–(N2)–Ae or Ae–(C6H6)–Ae bonds facilitate 2e loss but they are also weaker and more dynamic which impedes complex stability by ligand exchange reactions.16,18 Complexes with the larger Ae metals need to be stabilised by superbulky β-diketiminate (BDI) ligands.9,21 The replacement of the DIPP-substituents (DIPP = 2,6-iPr-phenyl) in the popular DIPPBDI ligand for much bulkier DIPeP-moieties in DIPePBDI (DIPeP = 2,6-CHEt2-phenyl) allowed for the synthesis of the first Ca dinitrogen complex.9 The additional advantage of the DIPeP-substituent is its solubilizing property which makes the final complexes even soluble in inert alkane solvents. Although successful in Ca chemistry, the same DIPePBDI ligand was insufficiently bulky for stabilisation of a similar Sr dinitrogen complex.16 However, using a dianionic bis-amide ligand with DIPeP-substituents enabled the isolation of a first Sr dinitrogen complex.16 The two additional K+ cations generate together with two Sr2+ cations a metal crown that efficiently encapsulates the N22− dianion (IV). Such complexes are reminiscent of K-aluminyl chemistry22 and Mulvey's inverse crown ethers.23

To our knowledge, highly reducing BaI synthons are hitherto non-existent. We recently reported a Ba–anthracene complex but considerable electron delocalisation in the anthracene dianion makes such complexes much more stable and less reducing.24 We now turn our attention to the challenging synthesis of a highly reactive Ba benzene complex. Complexes with this largest group 2 metal need stabilisation by our most bulky BDI ligand: DIPePBDI*. In this ligand, which also has been used in the synthesis of Mg0 complexes (II, III),6,8 the Me-backbone substituents have been replaced for large tBu-groups that direct the DIPeP rings towards the metal centre.25 Herein, we present a hybrid approach combining solid-state mechanochemical and solution methods to access Ba benzene inverse sandwich complexes. Completing the series of (BDI)Ae–(C6H6)–Ae(BDI) complexes (Ae = Mg, Ca, Sr, Ba), this now allows for discussion of trends in structure, electronics and reactivity.

Results and discussion

Syntheses and solution studies

Following a salt-metathesis protocol, a series of Ae iodide precursors [(DIPePBDI*)Ae(μ-I)]2 (Ae = Ca, Sr, Ba) was prepared from (DIPePBDI*)K and the corresponding AeI2 in yields of 92%, 46% and 87%, respectively (Scheme 2a and Fig. S6–S24, S97–S100). The potassium precursor was obtained in quantitative yield by deprotonation of DIPePBDI*–H with benzyl potassium and crystallised as the solvent-free cyclic tetramer [(DIPePBDI*)K]4 (SI: Fig. S1–S5 and S99).
image file: d5sc05373k-s2.tif
Scheme 2 (a) Syntheses of the precursors [(DIPePBDI*)Ae(μ-I)]2 (Ae = Ca, Sr, Ba). (b) Synthesis of Ae benzene inverse sandwich complexes by classical reduction in solution (1-Ca, 1-Sr). (c) Synthesis of Ba benzene inverse sandwich complexes by a hybrid solid-state/solution synthesis (1-Ba, 2-Ba).

Reduction of [(DIPePBDI*)Ae(μ-I)]2 (Ae = Ca, Sr) with 5% w/w K/KI26 or KC8 in benzene (Scheme 2b) led to the target compounds [(DIPePBDI*)Ae]266-C6H6) (1-Ae, Ae = Ca or Sr) in reasonable yields (1-Ca: 65%, 1-Sr: 42%). These yields are high when compared to those for similar complexes with the less bulky DIPePBDI ligand (Ca: 35%, Sr: 26%).9,18 The better yields for 1-Ae are due to more efficient steric protection and the lower solubility facilitating crystallisation.

The reduction of [(DIPePBDI*)Ba(μ-I)]2 with K/KI in benzene was found to be more problematic. Within two hours at room temperature, the Ba benzene inverse sandwich [(DIPePBDI*)Ba]266-C6H6) (1-Ba) had been formed selectively. However, the complex is highly soluble and upon removal of the benzene solvent under vacuum, it decomposed into a mixture in which mainly homoleptic species (DIPePBDI*)2Ba and small amounts of biphenyl were identified (Fig. S97). A similar phenomenon has been observed previously for (DIPePBDI)Sr–(C6H6)–Sr(DIPePBDI) (V). Highly concentrated benzene solutions led to biphenyl formation which could only be suppressed at low temperature.18 However, switching to the even more reactive Ba complexes, even concentration of the benzene solution by freeze-drying at −15 °C did not prevent product decomposition. Using an apolar aliphatic solvent with a small amount of benzene could enable product isolation by crystallisation, circumventing the need for solvent removal. However, this resulted in poor product selectivity. Lowering the temperature to 0 °C increased the selectivity, but drastically elongated reaction times.

Since a ball-milling approach has previously been very successful in Mg and Ca low-oxidation-state chemistry,18,29,30 we tried a hybrid solid-state/solution synthesis. The optimised protocol is as follows: [(DIPePBDI*)Ba(μ-I)]2 was ball-milled with a 4-fold excess of 5% K/KI for 30 minutes. After addition of a small amount of benzene, the resulting powder changed its colour from light-green to black, indicating formation of the C6H62− dianion. Subsequently, the product was extracted with n-pentane. Cooling the pentane solution to −35 °C led to crystallisation of [(DIPePBDI*)Ba]266-C6H6) (1-Ba) in 34% yield (Scheme 2c). This procedure also proved effective for reduction of the known [(DIPPBDI*)Ba(μ-I)]2 precursor.27 Due to its lower solubility in hydrocarbons, the product [(DIPPBDI*)Ba]266-C6H6) (2-Ba) crystallised much faster than 1-Ba, resulting in a higher yield of 54%.

The new complexes 1-Ae (Ca, Sr, Ba) and 2-Ba extend the family of known Ae benzene inverse sandwich complexes (Table 1) which now also includes the heaviest, least stable and most reactive Ba complexes. This allows for a comprehensive evaluation of trends.

Table 1 Stability of complexes of Ae benzene inverse sandwich complexes (BDI)Ae–(C6H6)–Ae(BDI) with BDI ligands of various bulk. Shown are the times for full decomposition monitored by 1H NMR spectroscopy in cyclohexane at room temperature, unless stated otherwise
Ae

image file: d5sc05373k-u1.tif

image file: d5sc05373k-u2.tif

image file: d5sc05373k-u3.tif

image file: d5sc05373k-u4.tif

a In benzene at 60 °C. b In toluene at 120 °C. c Compound not isolated but generated in situ in benzene. d Compound not obtained.
Mg 12 ha,20 3 daysb,17 d d
Ca 30 minc,18 2 days d 20 days
Sr d 1 day d 8 days
Ba d d 13 days 15 days


NMR data for the new complexes are shown in the SI (Fig. S25–S44 and S79). In all cases, the BDI* signals could be fully assigned. However, dianionic C6H62− remained silent, also upon cooling to −80 °C. Also in 13C NMR spectra no signals for the benzene dianion were observed. This is even the case in complexes with 99% 13C-labeled benzene. It is of interest to note that for all Ca, Sr and Ba benzene complexes reported so far, we have never observed NMR signals corresponding to the bridging C6H62− dianion.9,18 As discussed previously, this is likely due to a narrow singlet-triplet gap which allows thermal population of the open-shell state causing paramagnetic behaviour (see SI for details). Preliminary electron paramagnetic resonance (EPR) spectra of 13C6H6-bridged complexes also support the assignments of benzene dianions with triplet character (Fig. S108). In contrast, both Mg benzene inverse sandwich complexes show signals for the benzene dianion, which is likely due to their different nature. Whereas Ca, Sr and Ba complexes are highly ionic, Mg complexes have a significant covalent bonding component, resulting in a metalla-norbornadiene-like structure (Scheme 1, V).

While Ae benzene inverse sandwich complexes are at room temperature stable in the solid state, in solution they show only limited lifetimes. Direct comparison (Table 1) shows that their stabilities generally decrease with increasing metal size or decreasing ligand bulk. The family of complexes includes four different ligands with either DIPP- or the bulkier DIPeP-substituents (DIPPBDI or DIPePBDI) and backbone Me- or tBu-substituents (BDI or BDI*). Although the latter backbone substituents are far away from the metal centre, they have a considerable influence on steric shielding by pushing the N–Ar groups towards the remote metal.27,28 Especially the BDI*-based inverse sandwich complexes persist for weeks in alkane solutions at room temperature. There is also a clear correlation between complex stability and the size of the Ae metal. Using small ligands like DIPPBDI or DIPePBDI, the Ca and Sr complexes are increasingly unstable and the Ba complex could not even be obtained. The minimal ligand size required to stabilise the Ae–[C6H6]–Ae core: DIPPBDI for Mg,17DIPePBDI for Ca, Sr,18 and DIPPBDI* for Ba. Only with the larger set of BDI* ligands we have been able to isolate Ba benzene inverse sandwich complexes. With the largest ligand, DIPePBDI*, complex stability decreases from Ca to Sr but the Ba complex shows a surprisingly high stability and only fully decomposed after 15 days.

As described elsewhere for decomposition of Ca and Sr benzene inverse sandwich complexes,9,18 also the Ba complexes 1-Ba and 2-Ba decompose by release of benzene and form a mixture of unknown species (Fig. S80–S82). The only identified products of decomposition are homoleptic Ba species. Some of the signals in 1H NMR spectra of decomposed 1-Ba and 2-Ba (Fig. S82 and S83) match those of homoleptic [(DIPePBDI*)2Ba] and [(DIPPBDI*)2Ba] (Fig. S63 and S69), which we prepared for comparison according to alternative synthetic routes (see SI for details). We were not able to isolate or identify any products related to reductive C–H bond activation in the ligand, as observed previously in the decomposition of Ca and Sr benzene inverse sandwich complexes.9,18 The stability of 1-Ba and 2-Ba in benzene solution will be described below in the context reactivity studies.

Crystal structures

The molecular structures of the BDI complexes 1-Ca, 1-Sr and 1-Ba (Fig. 1 and S104–S106) are similar to those of previously reported Ca, Sr sandwich complexes9,18 and also show similarities to corresponding Yb or Sm complexes of a similar build-up.31 They are all binuclear metal complexes with bidentate BDI ligands and η66-bridging benzene moieties. However, there are marked structural differences and trends.
image file: d5sc05373k-f1.tif
Fig. 1 (a) Crystal structure of 1-Ba; H atoms omitted for clarity. (b) Selected bond distances (Å) and angles (°) in 1-Ca (blue), 1-Sr (purple) and 1-Ba (red). The penetrating van der Waals radii around the Ae metals and Cipso indicate bonding secondary Ae⋯Cipso(Ar) π-interactions.

Although the BDI ligand is generally described as a bidentate ligand, the structure of 1-Ba shows significant distortions (Fig. 1a). The aryl rings in the DIPePBDI* ligand are clearly bend to the Ba metal. This is evident from the N–Cipso⋯Cpara angles which are ideally 180° but in 1-Ba takes on values ranging from 163.2(7)° to 172.2(2)° (average: 169.4°). This distortion is less extreme for 1-Sr (average: 171.5°) and 1-Ca (average: 172.2°). Also the Cβ–N–Cipso angles are clearly distorted from the idealised 120° angle and increase from 1-Ca (average: 128.5°) to 1-Sr (average: 130.7°) to 1-Ba (average: 131.1°). The two Ar rings span a cavity in which the metal is bound through additional Ae⋯C(Ar) π-interactions. The latter interactions gain importance with metal size: Ca < Sr < Ba. This is evident from the angles between the Ar planes decreasing down the group: 1-Ca (101.5°) > 1-Sr (95.1°) > 1-Ba (92.9°). These interactions are partially due to the bulky tBu-substituents in the ligand backbone, pushing the Ar ring towards the metal. For comparison, the same ligand with backbone Me-substituents, DIPePBDI, shows much larger pockets with following angles between the Ar planes: Mg 141.8°,17 Ca 122.8°, Sr 131.8°. This demonstrates that the BDI* ligands with small metal pockets are superior in stabilisation of larger metals like Ca, Sr and especially Ba. Any Ae⋯C(π) distance below 3.13 Å (Ca), 3.30 Å (Sr) or 3.48 Å (Ba) could be considered a bonding interaction.32 Especially in saturating the large electropositive Ba2+ cation, such secondary interactions play a pivotal role and they can even compete with Ba–THF coordination.33 This may explain why the Ba benzene inverse sandwich complexes 1-Ba and 2-Ba are unexpectedly stable towards thermal decomposition (vide supra, Table 1). Since 1-Ba and 2-Ba are equally stable, these secondary Ae⋯C(Ar) π-interactions seem to outweigh the shielding advantage of Et2CH- versus Me2CH-substituents.

Apart from differences in the BDI ligand geometry, there are also clear trends in Ae–(C6H6)–Ae bonding. Whereas the bridging C6H62− ring in the Mg complexes is severely puckered, it gradually flattens going the larger Ae metals (Fig. 2a). It can be quantified by the values of the C–C–C–C torsion angles in the benzene ring which range from to 46.3(3)° for the Mg complex down to 1.4(4)° for 1-Ba. Flattening of the bridging benzene dianion may be explained by the increasing ionicity of the Ae–(C6H6) bond combined with the increasing size and softness of the Ae2+ cation. The hard Mg2+ cation is significantly more polarizing than the soft Ba2+ cation, thus favouring more covalent Mg–C 2e bonds. In contrast, the large Ba2+ prefers interaction with a large and soft C6H62− dianion. The amount of ring puckering also influences Ae–C bond distances. The most extreme puckered benzene ring shows a wide range of Mg–C bond distances: 2.209(1)–2.883(3) Å, while Ba–C bond distances for the least puckered benzene ring are in a narrow range 2.918(3)–3.026(4) Å (Fig. 2a). Whereas aromatic benzene has a planar structure with delocalised C–C bonds of equal length, the bridging (η66)-C6H62− dianions in all structures show unequal C–C bonds and varying degrees of ring puckering (Fig. 2a). Their geometries can be interpreted as a quinoid structure (two short and four longer C–C bonds) or a bis-allyl structure (four short and two longer C–C bonds). A comprehensive computational study on Li–(C6H6)–Li explains the origin of these distortions.34 In its most symmetric D6h form, the C6H62− dianion is in a singlet state in which the two additional electrons reside mainly in the periphery on the H atoms occupying a Rydberg molecular orbital (MO) with anti-bonding C–H character. A more stable alternative is the triplet state with single occupation of the two degenerated e2u orbitals (Fig. 2b), resulting in a C6H62− dianion with equal C–C bonds. However, the benzene dianion can be considerably stabilised by Jahn–Teller distortions. This either leads to a quinoid structure, resulting in lowering of the b1u MO, or to a bis-allyl structure which gives energy gain due to lowering of the au MO. Additionally, the second-order Jahn–Teller effect leads to a loss of [C6H6]2− planarity, which is gradually neutralised for heavier Ae2+ cations (Ca < Sr < Ba), which prefer ionic bonding with diffuse p-orbitals (Fig. 2a). In the crystal structures presented herein, the geometries of [C6H6]2− dianions point to a singlet ground state with either quinoid or bis-allyl distortions. However, as mentioned (vide supra), a narrow singlet-triplet gap allows thermal population of the open-shell state causing paramagnetic behaviour in solution. The recently reported Eu complex, [(DIPePBDI)Eu·THF]266-C6H6),35 showed the successful stabilisation of a long sought-after benzene dianion in the planar triplet state. This was enabled by antiferromagnetic coupling of two EuII nuclei through strong d–π* bonding interactions with the bridging [C6H6]2− dianion.


image file: d5sc05373k-f2.tif
Fig. 2 (a) Structures and selected geometric parameters for the Ae–(C6H6)–Ae moieties in Ae benzene inverse sandwich complexes stabilised by DIPePBDI or DIPePBDI* ligands (bond lengths/distances in Å and angles in °). The Ae–C and Ae–centroid distances are marked in red (for DIPePBDI* complexes) or in orange (for DIPePBDI complexes). The C–C bond lengths are shown in blue. The largest C–C–C–C torsion angles are marked in green. (b) Molecular orbital diagram for benzene and the effect of 2e reduction and Jahn–Teller distortions (1st and 2nd).

Computational studies

Molecular structures of 1-Ca, 1-Sr and 1-Ba were optimised in singlet closed-shell, singlet open-shell, and triplet states at the (U)B3PW91/def2-TZVP level of theory (see SI for details). To address the trends in the full series of Ae benzene inverse sandwich complexes, we included the structure of [(DIPPBDI)Mg]224-C6H6) which we abbreviate as 1′-Mg.17

Optimisation of all structures in the closed shell singlet state reproduced the crystal structures quite well (Tables S8 and S9). The dominant ionic character in all of these complexes is confirmed by relatively high positive NPA charges on the metals (1′-Mg: +1.69/+1.75, 1-Ca: +1.71/+1.74, 1-Sr: +1.74/+1.76, 1-Ba: +1.72/+1.75) and high negative charges on the bridging benzene dianions (−1.63/−1.56/−1.59/−1.58). The benzene dianions in the optimised structures adopt quinoidal structures, which is in agreement with crystal structures of 1′-Mg and 1-Ba, but contrasts with those of 1-Ca and 1-Sr, where bis-allyl forms are observed (Fig. 2a). This is in line with the very small energy difference between both isomers.34 The calculated structures generally present comparable ring puckering, when compared with experimental data. Calculated C–C–C–C torsion angles have values ranging from 47.1° (1′-Mg) to 3.2° (1-Sr and 1-Ba), while average C–C bond distances are in the 1.437–1.447 Å range (Fig. 3). NBO analysis shows that Mg–benzene bonding in 1′-Mg is comprised of Mg 3s-orbital and benzene π*-orbitals (b1u). The extreme boat form of the bridging benzene dianion results in asymmetric bonding. For the heavier Ae metals the ring flattens and not only valence s-orbitals but also empty d-orbitals participate in Ae–benzene bonding by accepting electron density (see Table S7). This is well presented by the calculated HOMO orbitals, which are formed by the overlap of the ns or ns/(n − 1)d orbitals of the Ae metal centres with the six 2pz atomic orbitals of the benzene carbon atoms (Fig. 3).


image file: d5sc05373k-f3.tif
Fig. 3 HOMO (isosurface 0.03 a.u.) representing Ae–(C6H6)–Ae bonding calculated for 1′-Mg, 1-Ca, 1-Sr and 1-Ba in the singlet state. Also shown is the geometry of the benzene dianion with C–C bond lengths (Å) marked in blue and the largest C–C–C–C torsion angles (°) marked in green.

Open-shell singlet/triplet calculations confirm that diradical character of the heavier 1-Ae complexes (Ca, Sr, Ba) is related to stable triplet states. These have been calculated to be 6.5 kcal mol−1 more stable for 1-Ca and are slightly more stable for 1-Sr and 1-Ba (1.3 kcal mol−1). In the triplet states, the benzene dianions show high D6h symmetry and are nearly planar (the largest C–C–C–C torsion angles: 0.6° for 1-Ca, 0.7° 1-Sr and 2.0° for 1-Ba). The C–C bond lengths are almost equal with values ranging from 1.433–1.437 Å (Fig. S109). In all cases, the SOMO and SOMO−1 in the triplet states are equivalent to the LUMO and HOMO in the singlet states, respectively (Fig. S110–S112).

Reactivity studies

Preliminary studies on the reactivity of Ba benzene inverse sandwich complexes are summarised in Scheme 3.
image file: d5sc05373k-s3.tif
Scheme 3 Reactivity of Ba benzene complexes depending on aryl (Ar = DIPP, DIPeP) substituents of ArBDI* ligand.

H2 activation

The methylcyclohexane solutions of 1-Ba and 2-Ba reacted instantaneously with 1.5 bar H2 at −120 °C. For 1-Ba the formation of the expected Ba hydride was confirmed by 1H NMR spectroscopy (Fig. S85). Diagnostic for a hydride ligand bound to Ba is the low-field singlet at 8.98 ppm, which is within the characteristic range of chemical shifts for Ba hydrides (7.92–10.39 ppm).36–38 However, although NMR spectroscopy confirmed hydride formation, the complex decomposed at room temperature over the course of two hours via ligand scrambling to homoleptic (DIPePBDI*)2Ba and insoluble BaH2 (Fig. S86). Any attempt to stabilize and crystallize the heteroleptic Ba hydride complex at −35 °C failed. Also the addition of O-donors like THF or THP did not lead to product crystallisation. Complex 2-Ba reacted fast with H2 but due to the less bulky DIPPBDI* ligand, only the well-characterised homoleptic complex (DIPPBDI*)2Ba was observed. Note, that the related Ba hydride complex [(DIPP/TCHPBDI)BaH]2 with the bulky DIPP/TCHPBDI ligand (HC{MeCN(DIPP)}{MeCN(TCHP)}, TCHP = 2,4,6-tricyclohexylphenyl) could be isolated and characterised in the solid-state. This is likely due to its poor solubility and rapid crystallisation within 10 minutes from the reaction mixture.39 However, this complex decomposed immediately upon dissolving in THF-d8. Good solubility and slow crystallisation of our complexes resulted in product decomposition. It should be noted that bulkier ligands not always work in favour of complex stability.40 Both homoleptic complexes, (DIPPBDI*)2Ba and (DIPePBDI*)2Ba, are spectroscopically characterised as asymmetric (κ2-BDI*)Ba(κ1-BDI*) complexes and resemble the previously reported (DIPePBDI)2Ca complex featuring a bidentate and monodentate BDI ligand.9 Whereas BDI ligands normally prefer bidentate (BDI)–Ae coordination,41 using superbulky ligands requires loss of one of the Ae–N bonds to relieve steric strain. For a symmetric (DIPPBDI*)2Ba complex with two bidentate ligands, we calculated that cleavage of one of the Ba–N bonds is highly exergonic: ΔG298 K = −37.6 kcal mol−1, ΔH0 = −52.8 kcal mol−1 (Table S5 and Fig. S113).

Reduction with I2 or (BDI*)Ae iodides

Reaction of 1-Ba and 2-Ba with iodine led to almost selective formation of the (BDI*)BaI precursors that were used in their synthesis (Fig. S87 and S88). Both Ba benzene complexes also reacted with lighter Ae iodide precursors: [(DIPePBDI*)AeI]2 (Ae = Mg, Ca, Sr). In case of Mg, benzene was eliminated and with moderate selectivity the formation of the MgI complexes [(DIPPBDI*)Mg]2 and [(DIPePBDI*)Mg]2 was observed (Fig. S89 and S90). This shows that 1-Ba and 2-Ba can substitute K0 and act as soluble reducing agents for the reduction of MgII to MgI. However, the anti-aromatic benzene dianion in these BaI synthons is not able to reduce CaII or SrII. Reaction of 1-Ba with [(DIPePBDI*)AeI]2 (Ae = Ca, Sr) in benzene or in cyclohexane did not result in metal reduction. Instead, we found an iodide/benzene ligand exchange, resulting in formation of 1-Ca and 1-Sr (Fig. S91 and S92). Interestingly, these ligand exchange processes are not reversible, i.e.1-Ca and 1-Sr do not react with [(DIPePBDI*)BaI]2. Therefore, the driving force for these ligand exchange processes must lie in differences in Ae–benzene and Ae–I bond strengths. Apparently, the Ba2+ cation has a very high affinity for the iodide anion.

Reduction of polyaromatic hydrocarbons

The reduction of selected polyaromatic hydrocarbons (PAHs) with 2-Ba is clean and selective as has been demonstrated for pyrene, biphenyl and 1,3,5-triphenyl-benzene (TPBz). In contrast, reaction of the bulkier complex 1-Ba led to formation of homoleptic (DIPePBDI*)2Ba as the main product. A possible explanation must be related to the only difference between both complexes: ligand bulk. Apparently, the extreme bulk of the DIPePBDI* ligand stabilizes the benzene complex 1-Ba but is not favourable for stabilisation of complexes with larger PAH2− dianions. We assume that after initial formation of a (DIPePBDI*)Ba–(PAH)–Ba(DIPePBDI*) complex, the product rapidly decomposes to (DIPePBDI*)2Ba and Ba(PAH). Such a scenario was found for reactions of 1-Ba with all PAHs that were tested (pyrene, biphenyl and TPBz). Note that this procedure enabled the isolation of (DIPePBDI*)2Ba, which could not be obtained selectively via a standard salt metathesis method (SI: Fig. S84, S93 and S94). Additional confirmation for decomposition of (DIPePBDI*)Ba–(PAH)–Ba(DIPePBDI*) species by the Schlenk equilibrium comes from the reaction of 1-Ba with 9,10-bis(trimethylsilyl)anthracene (TMS-Anth). Besides formation of (DIPePBDI*)2Ba, the previously reported complex [Ba(TMS-Anth)]n (ref. 24) was identified in the 1H NMR spectrum (Fig. S95).

The DIPPBDI*-based Ba inverse sandwich compounds 3-, 4- and 5-Ba containing pyrene, biphenyl, and TPBz dianions were prepared from in situ generated 1-Ba (Scheme 2c) and isolated as crystals in yields of 65%, 53% and 43%, respectively. Although all were isolated in crystalline form, crystallographic characterisation was only successful for the pyrene complex 3-Ba (Fig. S107). All Ba–PAH compounds present well-resolved NMR spectra, facilitating the complete assignment of 1H and 13C NMR signals (Fig. S45–S62). Though not crystallographically characterised, the most interesting complex is 5-Ba, for which NMR analysis points to η6-coordination of Ba centres to the inner ring of triphenylbenzene, in a similar manner like for Westerhausen's Ca inverse sandwich.42 This stands in contrast with a previously published Mg complex in which the (1,3,5-Ph3C6H3)2− dianion is sandwiched between two [(DIPPBDI)Mg]+ cations coordinating to one of its outer Ph rings (Fig. S78).30 This arrangement is likely dictated by steric congestion. Latter Mg compound was prone to decomposition by rearomatisation of triphenylbenzene and MgII → MgI reduction.

[C6H6]2− as nucleophile

The nucleophilic behaviour of the [C6H6]2− dianion in 1-Ba and 2-Ba is confirmed by its reactivity with iodobenzene, forming free biphenyl and (BDI*)Ba iodide by nucleophilic aromatic substitution (Fig. S96). Dissolved in benzene, both Ba benzene complexes also mediate dehydrogenative benzene–benzene coupling, which reduces their lifetimes in benzene solution to 2–3 days. Attempting to remove benzene solvent by freeze-drying at −15 °C led to biphenyl formation. In contrast to previously reported reactivity with Ca or Sr benzene complexes,18 the product is detected as neutral free biphenyl accompanied with homoleptic (BDI*)2Ba and presumably BaH2 (Fig. S97 and S98).

Conclusions

In order to isolate the most reactive and unstable Ae benzene inverse sandwich complex, (BDI)Ba–(C6H6)–Ba(BDI), we used a combination of ball-milling and solution chemistry. To stabilize such labile complexes, superbulky BDI ligands are crucial.

The availability of benzene inverse sandwich complexes throughout the series Mg, Ca, Sr and Ba allows for evaluation of metal effects on stability, structure, electronics and reactivity. Using the DIPePBDI ligand allowed for isolation benzene inverse sandwich complexes for the metals Mg, Ca and Sr which show increased lability with metal size. For the bigger ligand DIPePBDI* with tBu-backbone substituents, Ca, Sr and Ba complexes were isolated. The crucial ligand architecture required for the challenging isolation of the Ba complexes are the tBu-substituents in the ligand backbone. These push the Ar rings towards the large Ba metal cations, stabilizing the complexes with secondary Ba⋯(π-Ar) interactions. The latter results in an unexpected high stability for these Ba complexes which dissolved in alkanes decompose over a period of up to 15 days.

Another noticeable trend is found in the structure of the bridging benzene dianions. Whereas in the Mg–(C6H6)–Mg moiety the ring is strongly distorted from planarity and features a boat conformation, moving to heavier Ae metals result in flattening of this structure. The ring in the Ba–(C6H6)–Ba moiety is essentially flat. All crystal structures show rings with combinations of short and long C–C bonds as is typical for closed-shell singlet states. This can either be a quinoidal structure (two short C[double bond, length as m-dash]C bonds and four longer C–C bonds) or a bis-allyl structure (two long C–C single bonds and four shorter allylic C–C bonds). Interestingly, in solution all benzene rings are NMR silent due to paramagnetic behaviour. This is due to low-lying triplet states in which the benzene ring is planar and the two unpaired electrons are in degenerate e2u orbitals resulting in equal C–C bond distances. DFT calculations predict hardly any energy differences between these states. NBO analysis shows that all complexes have a large extent of ionic character with high positive charges on the metals (+1.69/+1.76) and high negative charges on the benzene ligands (−1.58/−1.63).

Preliminary reactivity studies of inverse sandwich complexes of type (BDI)Ba–(C6H6)–Ba(BDI) show fast reaction with H2 to give (BDI)BaH and benzene, showing that they are efficient BaI synthons. Unfortunately, while (DIPePBDI*)BaH could be detected in solution, its high instability precluded further characterisation by crystal structure determination. Heteroleptic (BDI)BaH complexes decomposed in solution to homoleptic products (BDI)2Ba and BaH2. The complex (DIPePBDI*)2Ba shows the typical NMR signals for an asymmetric complex with one bidentate and one mono-dentate BDI ligand. This is a demonstration of the bulk of the DIPePBDI* ligand. Even for the largest group 2 metal Ba, a complex with two bidentate ligands was calculated to be 37.6 kcal mol−1G298 K) higher in energy than the bidendate/monodentate combination.

We could also demonstrate that these BaI synthons The BaI synthons are able to reduce a range of PAH's. However, more interesting is the MgII to MgI reduction. Reaction of (BDI)Ba–(C6H6)–Ba(BDI) with a (BDI)MgI precursor led to formation of (BDI)Mg–Mg(BDI), (BDI)BaI and benzene. However, heavier Ae metal cations like Ca2+ or Sr2+ could not be reduced. Reaction of (BDI)Ba–(C6H6)–Ba(BDI) reagents with a (BDI)AeI (Ae = Ca, Sr) precursor led to formation of the corresponding (BDI)BaI and (BDI)Ae–(C6H6)–Ae(BDI). This iodide/benzene2− exchange does not work in the other direction, showing that the lighter benzene inverse sandwich complexes are more stable than the heavier ones.

Finally, the benzene2− dianion in (BDI)Ba–(C6H6)–Ba(BDI) can also act as a nucleophile. It is well-known that the heavier Ae metals can facilitate nucleophilic aromatic substitution. This is for example key to H/D exchange at aromatic rings using Ae metal hydride catalysts.43 Thus, reaction of (BDI)Ba–(C6H6)–Ba(BDI) with PhI or simply benzene led to biphenyl formation.

We are currently exploring further reactivity of these first BaI synthons.

Author contributions

D. Jędrzkiewicz: conceptualisation, investigation, validation, formal analysis, writing – original draft, visualisation. M. Morasch, O. P. E. Townrow, B. Rösch: investigation, validation, formal analysis. J. Langer: formal analysis, validation. Z. Mathe: formal analysis, validation. S. Harder: conceptualisation, writing – original draft – review and editing, validation, supervision, project administration.

Conflicts of interest

There are no conflicts to declare.

Data availability

CCDC 2435047–2435054 and 2435056 contain the supplementary crystallographic data for this paper.44a–i

Supplementary information: Synthetic procedures, selected NMR spectra, details for crystal structure determination, details for EPR characterization, details for computational work. See DOI: https://doi.org/10.1039/d5sc05373k.

Acknowledgements

We acknowledge Mrs A. Roth (University of Erlangen-Nürnberg) for CHN analyses, J. Schmidt and Dr C. Färber (University of Erlangen-Nürnberg) for assistance with the NMR analyses and L. Klerner for assistance with NPA analysis. S. H thanks the Deutsche Forschungsgemeinschaft for funding (HA 3218/11). O. P. E. T. thanks the Alexander von Humboldt Foundation for a postdoctoral fellowship, VCI for a Liebig fellowship and support by the state of Baden-Württemberg through bwHPC and the Deutsche Forschungsgemeinschaft (DFG) through grant no. INST 40/575-1 FUGG for access to the JUSTUS 2 cluster for computational work. Z. M. thanks the Max Planck Society for funding. We are grateful to Johannes Maurer and Stefan Thum for optimizing the synthetic route to 2,6-di(3-pentyl)aniline.

References

  1. S. P. Green, C. Jones and A. Stasch, Science, 2007, 318, 1754 CrossRef CAS.
  2. S. P. Green, C. Jones and A. Stasch, Angew. Chem., Int. Ed., 2008, 47, 9079 CrossRef CAS.
  3. S. J. Bonyhady, C. Jones, S. Nembenna, A. Stasch, A. J. Edwards and G. J. McIntyre, Chem.–Eur. J., 2010, 16, 938 CrossRef CAS.
  4. C. Jones, Nat. Rev. Chem., 2017, 1, 0059 CrossRef CAS.
  5. J. T. Boronski, A. E. Crumpton, L. L. Wales and S. Aldridge, Science, 2023, 380, 1147 CrossRef CAS PubMed.
  6. B. Rösch, T. X. Gentner, J. Eyselein, J. Langer, H. Elsen and S. Harder, Nature, 2021, 592, 717 CrossRef.
  7. C. Berthold, J. Maurer, L. Klerner, S. Harder and M. R. Buchner, Angew. Chem., Int. Ed., 2024, e202408422 CAS.
  8. J. Mai, J. Maurer, J. Langer and S. Harder, Nat. Synth., 2024, 3, 368 CrossRef CAS.
  9. B. Rösch, T. X. Gentner, J. Langer, C. Farber, J. Eyselein, L. Zhao, C. Ding, G. Frenking and S. Harder, Science, 2021, 371, 1125 CrossRef.
  10. J. Mai, B. Rösch, N. Patel, J. Langer and S. Harder, Chem. Sci., 2023, 14, 4724 Search PubMed.
  11. B. Maitland, A. Stasch and C. Jones, Aust. J. Chem., 2022, 75, 543 Search PubMed.
  12. Y. Xie, H. F. Schaefer III and E. D. Jemmis, Chem. Phys. Lett., 2005, 402, 414 CrossRef CAS.
  13. M. A. Gosch and D. J. D. Wilson, Organometallics, 2023, 42, 2185 CrossRef CAS.
  14. R. Mondal, K. Yuvaraj, T. Rajeshkumar, L. Maron and C. Jones, Chem. Commun., 2022, 58, 12665 RSC.
  15. R. Mondal, M. J. Evans, T. Rajeshkumar, L. Maron and C. Jones, Angew. Chem., Int. Ed., 2023, 62, e202308347 CrossRef CAS PubMed.
  16. M. Morasch, T. Vilpas, N. Patel, J. Maurer, S. Thum, M. A. Schmidt, J. Langer and S. Harder, Angew. Chem., Int. Ed., 2025, e202506989 CAS.
  17. T. X. Gentner, B. Rösch, G. Ballmann, J. Langer, H. Elsen and S. Harder, Angew. Chem., Int. Ed., 2019, 58, 607 CrossRef CAS PubMed.
  18. J. Mai, M. Morasch, D. Jędrzkiewicz, J. Langer, B. Rösch and S. Harder, Angew. Chem., Int. Ed., 2023, 62, e202212463 CrossRef CAS.
  19. J. Mai, B. Rösch, J. Langer, S. Grams, M. Morasch and S. Harder, Eur. J. Inorg. Chem., 2023, 26, e202300421 CrossRef CAS.
  20. D. D. L. Jones, I. Douair, L. Maron and C. Jones, Angew. Chem., Int. Ed., 2021, 60, 7087 CrossRef CAS PubMed.
  21. 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 CrossRef CAS.
  22. J. Hicks, P. Vasko, J. M. Goicoechea and S. Aldridge, Angew. Chem., Int. Ed., 2020, 60, 1702 CrossRef PubMed.
  23. R. E. Mulvey, Organometallics, 2006, 25, 1060 CrossRef CAS.
  24. O. P. E. Townrow, C. Färber, U. Zenneck and S. Harder, Angew. Chem., Int. Ed., 2024, 63, e202318428 CrossRef CAS PubMed.
  25. B. Rösch, T. X. Gentner, J. Eyselein, A. Friedrich, J. Langer and S. Harder, Chem. Commun., 2020, 56, 11402 RSC.
  26. J. Hicks, M. Juckel, A. Paparo, D. Dange and C. Jones, Organometallics, 2018, 37, 4810 CrossRef CAS.
  27. H. M. El-Kaderi, M. J. Heeg and C. H. Winter, Polyhedron, 2006, 2, 224 CrossRef.
  28. C. Chen, S. M. Bellows and P. L. Holland, Dalton Trans., 2015, 44, 16654 RSC.
  29. D. Jędrzkiewicz, J. Mai, J. Langer, Z. Mathe, N. Patel, S. DeBeer and S. Harder, Angew. Chem., Int. Ed., 2022, 61, e202200511 CrossRef PubMed.
  30. D. Jędrzkiewicz, J. Langer and S. Harder, Z. Anorg. Allg. Chem., 2022, 648, e202200138 CrossRef.
  31. S. Kumar Thakur, N. Roig, R. Monreal-Corona, J. Langer, M. Alonso and S. Harder, Angew. Chem., Int. Ed., 2024, 63, e202405229 CrossRef.
  32. W. D. Buchanan, D. G. Allis and K. Ruhlandt-Senge, Chem. Commun., 2010, 46, 4449 RSC.
  33. P. M. Chapple and Y. Sarazin, Eur. J. Inorg. Chem., 2020, 35, 3321 CrossRef.
  34. A. Falceto, D. Casanova, P. Alemany and S. Alvarez, Chem.–Eur. J., 2014, 20, 1 CrossRef PubMed.
  35. Y. Wang, R. Sun, J. Liang, Y. Zhang, B. Tan, C. Deng, Y.-H. Wang, B.-W. Wang, S. Gao and W. Huang, J. Am. Chem. Soc., 2025, 147(9), 7741 CrossRef CAS PubMed.
  36. M. Wiesinger, B. Maitland, C. Färber, G. Ballmann, C. Fischer, H. Elsen and S. Harder, Angew. Chem., Int. Ed., 2017, 56, 16654 CrossRef CAS PubMed.
  37. X. Shi, C. Hou, C. Zhou, Y. Song and J. Cheng, Angew. Chem., Int. Ed., 2017, 56, 16650 CrossRef CAS PubMed.
  38. X. Shi, G. Qin, Y. Wang, L. Zhao, Z. Liu and J. Cheng, Angew. Chem., Int. Ed., 2019, 58, 4356 CrossRef CAS PubMed.
  39. D. B. Kennedy, M. J. Evans, D. D. L. Jones, J. M. Parr, M. S. Hill and C. Jones, Chem. Commun., 2024, 60, 10894 RSC.
  40. M. Arrowsmith, B. Maitland, G. Kociok-Köhn, A. Stasch, C. Jones and M. S. Hill, Inorg. Chem., 2014, 53, 10543 CrossRef CAS PubMed.
  41. S. Harder, Organometallics, 2002, 21, 3782 Search PubMed.
  42. S. Krieck, H. Görls, L. Yu, M. Reiher and M. Westerhausen, J. Am. Chem. Soc., 2009, 131, 2977 CrossRef CAS.
  43. J. Martin, J. Eyselein, S. Grams and S. Harder, ACS Catal., 2020, 10(14), 7792 Search PubMed.
  44. (a) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435047: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqvx4; (b) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435048: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqvy5; (c) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435049: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqvz6; (d) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435050: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqw08; (e) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435051: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqw19; (f) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435052: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqw2b; (g) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435053: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqw3c; (h) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435054: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqw4d; (i) D. Jędrzkiewicz, M. Morasch, O. P. E. Townrow, B. Rösch, J. Langer, Z. Mathe and S. Harder, CCDC 2435056: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mqw6g.

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