Róbert Csonka,
Gábor Speier and
József Kaizer*
Department of Chemistry, University of Pannonia, 8201 Veszprém, Hungary. E-mail: kaizer@almos.vein.hu
First published on 4th February 2015
During the past decade isoindoline-based ligands became the subject of growing interest due to their modular set-up. In this review the structure and reactivity of these ligands and their transition metal complexes are covered. Beyond the discussion of the structural properties particular attention is paid to the expanding fields of applications of these compounds.
The main structural features involve bis(arylimino), bis(alkylimino) and monosubstituted (asymmetric) isoindolines. Another structure of interest belongs to the phthalazine type molecules originated from bis(R-imino)isoindolines (BII or 3) by ring expansion with hydrazine hydrate.14 The transition metal complexes of these formulations have applications from enzyme mimics15–17 to catalytic hydrogenation18 or in photophysics,19 that is discussed in this review.
By the presence of aliphatic moieties on BII the chelating capacity is usually limited. The lack of heteroatoms favors the mono-, or bidentate ligand formation that is based on the coordination to the endocyclic NH and one imino side arm on the isoindole ring. Although, aliBIIs have been reported since the first appearance of BIIs,2,20,26–31,54 complexes of this kind have been synthesized only once by Maleev et al. In this work, the coordination of bidentate aliBII to rare earth metals is discussed. They are observed both in terminal and bridge positions using all three coordination sites in the latter case.32
The less common monodentate behavior of BII has been detected by our group33 by blocking the N atoms of 1,3-bis(2′-pyridylimino)isoindoline – BPI (7). The zwitterionic copper(I) compound showed moderate catechol oxidase activity.
BPI and its substituted derivatives are the most researched aroBIIs. They are considered pincer type ligands based on the tridentate coordination mode and the aromatic planarity around the metal ions. They are monoanionic, nevertheless complexes of protonated ligands can form as well with suitable metal ions and the proper conditions applied.34 The meridional configuration is open enough to host other coligands or various substrates. A second identical ligand stabilizes the structure creating an octahedral, homoleptic complex, although the active species remains usually the 1:
1 ligand to metal ratio form.
Because of the favorable characteristics of BPI chiral side chains can be attached as well. Fletcher et al. reported a facile one pot synthesis of chiral aminopyridines.35 To obtain any optically active BII is only one step away. Myrtene, α-pinene and carene have been used to build chiBIIs. Concerning enantioselective catalysis only a few number of work can be found done by Bröring and Gade.13,36–38
The preparation of BIIs by all known methods is a two-step procedure. The partial condensation of the starting material might happen if reaction time is not adequate. This can be used to our advantage if the goal is to achieve unsyBII. Since intramolecular C–H bond activation has been observed multiple times by Pd complexes,39–41 a better understanding on the mechanism is greatly expected. In order to study this phenomenon several unsyBIIs have been synthesized.42,43 Unsymmetry have been proved to be useful in the research of catecholase activity as well.23
Versatility of BPI goes beyond its applications and side chain variations. Simply, by the addition of hydrazine hydrate a ring expansion occurs on the isoindoline core that reveals an entirely new chemistry (Scheme 2).14
PAP (4) has a similar behavior to 3. It is anionic and it can form any metal complexes both in protonated or deprotonated state. Acetato, hydroxo, oxo, chloro etc. bridges are often observed in the binuclear center. The number of coordinated anions represented by L on Scheme 2 can vary depending on the metal salt used.
In Fig. 2 ligands have been collected that are the derivatives of both 3 and 4 showing only the R-group and their reference number used in this review.
The second key structural element is the endocyclic NH group. It is responsible for an anionic character observed with metal ions. Depending on the electronic strength of R-groups, the anion applied with the metal ion, the solvent used, the inert atmosphere and the presence of bases both deprotonated and protonated complexes can be formed.
The electronic absorption spectrum of the ligands 3 show a multiple band pattern. The π–π* transitions are created by the extended π-bond system and they appear in the 360–500 nm region. Upon complex formation small (2–20 nm) red or blue shifts are observed. Lower energy bands are contributed to charge transfer transitions from metal ion to the ligand. A selection of absorption bands of various BII and their homoleptic transition metal complexes has been collected in Table 1.
Ligand | π–π* CT bands of the free ligands λmax/nm (log![]() |
[M(L)2]λmax/nm (log![]() |
Ref. |
---|---|---|---|
a Ligand synthesis.b Complex synthesis and X-ray.c Spectroscopic data. | |||
7 | 410 (3.99), 386 (4.23), 366 (4.19) in DMF | 440 (4.58), 410 (4.81), 390 (4.75) (FeII) | 1a, 48b and 22c |
8 | 408 (4.11), 385 (4.29), 366 (4.24) in DMF | 450 (4.65), 423 (4.73), 332 (4.45), 306 (4.53) (MnII) | 1a, 49b, 22c and 49c |
9 | 470 (4.01), 440 (4.30), 413 (4.31), 388 (4.35), 369 (4.37) in DMF | 493 (4.28), 453 (4.49), 389 (4.40), 367 (4.45) (FeII) | 50a, 22b,c and 51b,c |
474 (4.71), 440 (4.86), 414 (4.86) 364 (4.94), 287 (4.87) (FeIII) | |||
10 | 485 (3.18), 448 (3.86), 420 (4.05), 396 (4.01), 374 (3.94) in DMF | 476 (4.11), 433 (4.54), 410 (4.55), 333 (4.38), 319 (4.40), 286 (4.45) (CoII) | 52a, 53b, 22c and 53c |
![]() | ||
Fig. 3 A dimolybdenum complex (36) with BPI (7).56 |
As a remarkable feature of 36, one molybdenum ion is coordinated to an imino nitrogen rotating a pyridyl arm into a position where binding to the dimetal unit is not possible. Similar “imino binding” has been documented only once with an aliBII ligand32 [1,3-bis(isopropylimino)isoindolinate] but it could be expected in that case on the account of missing coordinating sidearms.
Table 2 contains the characteristic data of the manganese complexes of BIIs. The complexes are identified using the abbreviations reported in the original manuscripts with the corresponding R-group from Fig. 2, being denoted for reference with this review. Tau values are valid only for five coordinate structures. The numbers show a nearly perfect trigonal bipyramid (42) and an average, moderately distorted TBPY (38). The influential factors such as oxidative atmosphere, bases and the right anion that greatly determine the BII structure will be described in the section of Fe-group elements due to the greater amount of research data. The distance between the amino group of the isoindoline core and the metal ion varies in a narrow interval among the manganese complexes. The difference is prominent at +2 (45) and +3 (46) oxidation states when the same ligand coordinates that can be explained by Jahn–Teller effect. Fig. 4 also emphasizes this difference. Only the Mn2+ complexes fit to a linear tendency with the increasing Nind–Mn2+ distance.
Ligand | Complex | NPy–Mna (Å) | Nind–Mnb (Å) | τ | Ref. |
---|---|---|---|---|---|
a The average pyridyl N–Mn distance.b The isoindoline N(H)–Mn distance. | |||||
7 | [MnII(ind)2] (37) | 2.295 | 2.163, 2.143 | — | 49 |
7 | [MnII(indH)Cl2] (38) | 2.249 | 2.153 | 0.69 | 17 |
11 | [MnII(3′Me-BPI)2] (39) | 2.293 | 2.144, 2.151 | — | 21 |
8 | [MnII(4′Me-ind)2] (40) | — | — | — | 49 |
12 | [MnII(6′Me2indH)(H2O)2(CH3CN)](ClO4)2 (41) | — | — | — | 16 |
9 | [MnII(bimindH)Cl2](DMF) (42) | 1.959 | 2.007 | 0.93 | 50 |
9 | [MnII(bimind)2] (43) | — | — | — | 15 |
13 | [MnII(Mebimind)2] (44) | — | — | — | 15 |
10 | [MnII(BTI)2] (45) | 2.220 | 2.211, 2.220 | — | 15 |
10 | [MnIII(BTI)2] (46) | 2.146 | 1.968, 1.967 | — | 59 |
Cobalt and nickel complexes of the title ligands are fairly rare although they have been used multiple times in the oxidation, hydroxylation of hydrocarbons (see also chapter 4.2.).60,61 The “pincer” shape is also an optimal platform for enantioselective catalytic reactions.36 Interaction with DNA have been studied in order to develop DNA cleaving therapeutic agents and non-radioactive probes for DNA through Co(BPI)s.62
Ligand | Complex | NPy–Fea (Å) | Nind–Feb (Å) | τ (Φ) | Ref. |
---|---|---|---|---|---|
a The average pyridyl N–Fe distance.b The isoindoline N(H)–Fe distance. | |||||
7 | [FeII(ind)2] (47) | 2.240 | 2.069 | — | 48 |
7 | [FeII(ind)(CH3CN)3](ClO4)2 (48) | 2.200 | 2.072 | — | 34 |
7 | [FeIII(μ-O)(ind)2Cl2]THF (49) | 2.153 | 1.998 | 0.88 | 63 |
7 | [FeIII(ind)Cl2] (50) | 2.148 | 1.963 | 0.86 | 64 |
8 | [FeIII(4′Me-ind)Cl2] (51) | 2.144 | 1.978 | 0.77 | 65 |
9 | [FeIII(bimind)2] (52) | 1.979 | 1.912, 1.928 | — | 51 |
9 | [FeII(bimind)Cl2] (53) | 2.067 | 2.045 | — | 66 |
13 | [FeII(Mebimind)2] (54) | 2.136 | 2.057, 2.053 | — | 22 |
10 | [FeIII(BTI)2]MeCN (55) | 2.002 | 1.946, 1.952 | — | 22 |
10 | [FeIII(BTI)2] (56) | 1.982 | 1.928, 1.922 | — | 59 |
10 | [FeIII(BTI)Cl2] (57) | 2.095 | 2.019 | 0.83 | 65 |
14 | [FeIII(5′Me-BTI)Cl2] (58) | 2.098 | 2.029 | 0.59 | 67 |
15 | [FeIII(μ-O)(benzBTI)2Cl2] (59) | 2.165 | 2.064, 2.052 | 0.76 | 68 |
16 | [FeIICI(DMSO)(thqbpi)] (60) | 2.169 | 1.996 | 0.31 | 69 |
16 | [FeII(thqbpi)(CH2SiMe3)] (61) | 2.194 | 2.002 | 0.65 | 70 |
7 | [FeII(4-MeBPI)2] (62) | 2.085 | 1.979 | — | 61 |
A large number of transition metal complexes of BIIs have been reported by Robinson et al. in 1967.71 Unfortunately, at this time the lack of accurate analytical techniques limited the clear insight into fine structures. Nowadays, it is known that anions may participate in the inner and outer coordination sphere of the crystals. Similar roles have been assigned to solvent molecules too. The presence of water or non-inert atmosphere increases the chance of oxo-bridged formulations or the replacement of either anions or other solvents in the first coordination sphere (or inner sphere). Based on the available structures and synthetic procedures some conclusions can be made on complex formation. The iron(II)perchlorate and acetate salts tend to result in a 2:
1 (ligand–metal) complex (52, 54, 62, furthermore 37, 39 with manganese content) mainly with a help of some base (e.g. triethylamine). In the absence of any base and under inert atmosphere protonated 1
:
1 complex can be achieved in a good coordinating solvent such as acetonitrile (48). Chlorides seem to favor the TBPY-5 set. There are only 1
:
1 complexes reported both with iron(II) and iron(III)chlorides (50, 51, 58, similarly with Mn: 38, 42). Two precedents show that a μ-oxo bridge forms between two iron cores with FeCl3 (49, 59) under air.
In spite of the attempts to obtain FeII(BTI)2 two independent work proved the viability of FeIII(BTI)2 (55, 56) submitted almost at the same time.22,59 They are seemingly identical (if anions are ignored) but X-ray structures reveal more. If average Naromatic–Fe distances are plotted against NInd–Fe distances high- and low-spin complexes can be separated (Fig. 5). There is Mössbauer spectra available to support the high-spin nature of 47 and 54.22,48 In a work by Pap et al. a low-spin Fe3+ complex (52) has been synthesized from its high-spin Fe2+ counterpart.51 55 was obtained from the Fe2+ form as well but less harsh conditions were applied.22 At the same time 56 was prepared directly from a Fe3+ salt undoubtedly creating a high-spin complex.
The only ruthenium complexes of BII available in literature have been presented by Pitchumony et al.,72 Gagné et al.73,74 and Tseng et al.75 all together with three X-ray structures (63–65) (Fig. 6). Ligands 7, 8 and 12 (respectively) were involved in the investigations. The resolved structure in the first case (63) contains a ligand fracture (1,2-dicyanobenzene) impurity that has an unusual monodentate connection to the metal center. Ruthenium complexes of 8 have been synthesized in various forms, just as Ru(4′-MeLH)Cl3 and Ru(4′-MeL)2. They were identified by spectroscopic methods and electrochemical and catalytic oxidative properties were tested. Complexes of 12 have PPh3 coligands RuII(6′Me-BPI)(PPh3)Cl (64) and HRuII(6′Me-BPI) (PPh3)2 (65). They catalyze base-free and chemoselective alcohol dehydrogenation.
![]() | ||
Fig. 6 Representation of known ruthenium structures (63–65).72,75 |
Cobalt complexes of any kind favor the octahedral symmetry more than any other metals. The absence of the Werner-type arrangement is observed only in a couple of cases (72, 73, 75, 76, 78 – Table 4). 75–78 from these molecules have been intentionally converted to the TBPY-5 structure in order to investigate activity in radical polymerization.77 A rarely identified structure of 74 was achieved by pyridine displacement from the metal dialkyl precursor (py)2Co(CH2SiMe3)2.
Ligand | Complex | NPy–Coa (Å) | Nind–Cob (Å) | τ (Φ) | Ref. |
---|---|---|---|---|---|
a The average pyridyl N–Co distance.b The isoindoline N(H)–Co distance. | |||||
7 | [CoIII(ind)2](ClO4)MeOH (66) | 1.985 | 1.889 | — | 62 |
7 | [CoIII(BPI)(OBz)(OO-t-Bu)] (67) | 1.955 | 1.845 | — | 60 |
9 | [CoII(bimind)OAc DMF] (68) | 2.093 | 2.040 | — | 66 |
10 | [CoII(BTI)2] (69) | 2.137 | 2.069, 2.073 | — | 53 |
10 | [CoIII(BTI)2] (70) | 1.959 | 1.913, 1.931 | — | 59 |
11 | [CoII(3′Me-ind)2] (71) | 2.196 | 2.004, 2.005 | — | 76 |
16 | [CoIICl(Me2SO)(thqbpi)] (72) | 2.161 | 1.941 | 0.51 | 69 |
16 | [Co2II(OAc)(thqbpi)2(MeOH)(EtOH)] (73) | 2.148 | 1.941, 1.939 | 0.54 | 69 |
17 | [CoII(CH2SiMe3)(pentbpi)] (74) | 2.076 | 1.949 | 0.67 | 37 |
18 | [Co2II(dihexyl1-tBu-bpi)2(OAc)] (75) | 2.126 | 1.983, 1.975 | 0.71 | 77 |
19 | [CoII(acac)(5′chloro-bpi)] (76) | 2.159 | 1.969 | 0.64 | 77 |
19 | [CoII(acac)(3,4-dichloro-5′chloro-bpi)(MeOH)] (77) | 2.177 | 2.021 | — | 77 |
19 | [CoII(acac)(3,4-dichloro-5′chloro-bpi)] (78) | 2.156 | 1.972 | 0.50 | 77 |
The two other members – rhodium and iridium – of this group belong to the rarest elements of the Earth's crust. A paper by Gade's research group presented a BII complex of the second densest element. IrI(18) (COD) with cyclooctadiene coordination was successfully identified by X-ray resulting in a OC-6 structure.78 18 behaves as a bidentate ligand that was observed only at lanthanides32 and at aliBIIs. A greater number of carbonyl complexes of the corresponding metals with aroBIIs were reported by Siegl.79 These compounds were identified by IR and elemental analysis and their reactivity toward alkyl and acyl halides was tested.
Ligand | Complex | NPy–Nia (Å) | Nind–Nib (Å) | Φ | Ref. |
---|---|---|---|---|---|
a The average pyridyl N–Ni distance.b The isoindoline N(H)–Ni distance. | |||||
7 | [NiII(ind)2] (79) | 2.168 | 2.024 | — | 15 |
7 | [NiII(ind)(2-Clpcyd)] (80) | 1.965 | 1.827 | — | 80 |
8 | [NiII(4′Me-ind)2] (81) | 2.180 | 2.017 | — | 15 |
9 | [NiII(bimind)OAc DMF] (82) | 2.069 | 2.011 | — | 66 |
20 | [NiII(boxmi)Cl] (83) | 1.911 | 1.904 | 0.07 | 36 |
The enantioselective fluorination of oxindoles and β-ketoesters to the corresponding products is effectively catalyzed by 83 with high ee and noteworthy yield.36
The most redundant group of isoindoline complexes is undoubtedly the one with palladium core. Its popularity can be attributed to the cyclopalladation phenomena. Unlike with other transition metals, unique N,N,C type coordination can occur with BII that was documented by Dietrich et al.39 Steric hindrance on the pyridyl moiety and palladium excess are the influential factors of cyclometallation.40 The flexibility of BIIs allows a flip on the axis of one C–N bond that is attributed to an intramolecular strain.41 After C–H and C–C bond activation takes place further PdII can coordinate to exocyclic N atoms (87). Exploiting this remarkable feature of palladium S-coordination (114) was observed on one thiazole subunit of BTI ligand (26) that was unprecedented before (Fig. 7).86
![]() | ||
Fig. 7 Exocyclic- (left) and S-coordination (right) of palladium complexes.41,86 |
All Pd complexes are planar or pseudoplanar, no other structures have been recorded. Unsymmetrical complexes were not included in Table 6. Kleeberg and Bröring reported a series of such complexes with unusual structural features.43 The new complexes had tBu-selenazol-pyridine and tBu-thiazole-pyridine sidearms on their ligands. In both of their packing pattern large voids were comprised that could be suitable even for small molecule storage according to the authors.
Ligand | Complex | NPy–Pda (Å) | Nind–Pdb (Å) | Φ | Ref. |
---|---|---|---|---|---|
a The average pyridyl N–Pd distance.b The isoindoline N(H)–Pd distance. | |||||
7 | [PdII(BPI)(OAc)] (84) | 2.038 | 1.946 | 0.13 | 41 |
8 | [PdIICl(4-(HOCH2CH2)-10-MeBPI)] (85) | 2.063 | 1.965 | 0.09 | 81 |
12 | [PdII(6′Me-ind)(Cl)] THF (86) | 2.152 | 1.967 | 0.19 | 39 |
12 | [PdII4(6′Me-BPI)2(OAc)4] (87) | 2.174 | 1.987 | 0.16 | 41 |
18 | [PdIICl(4-Me-10-tBuBPI)] (88) | 2.065 | 1.957 | 0.18 | 18 |
21 | [PdIICl(11-Me-BPI)] (89) | 2.058 | 1.93, 1.94 | 0.09 | 18 |
22 | [PdIICl(11-Br-BPI)] (90) | 2.061 | 1.962 | 0.16 | 18 |
23 | [PdIICl(11-TMS-ethynyl-BPI)] (91) | 2.049 | 1.958 | 0.21 | 18 |
24 | [PdIICl(11-Ph3Si-ethynyl-BPI) (92) | 2.031 | 1.952 | 0.23 | 18 |
25 | [PdII(pinBPI)(OAc)] (93) | 2.053 | 1.936 | 0.04 | 38 |
25 | [PdIICl(pinBPI)] (94) | 2.053 | 1.962 | 0.20 | 38 |
26 | [PdII(4-MeBTI)(OC(NH2)Me)]+ BArF− (95) | 2.036 | 1.968 | 0.29 | 82 |
26 | [PdII(4-MeBTI)(dmf)]+ BArF− (96) | 2.026 | 1.968 | 0.24 | 82 |
26 | [PdII(4-MeBTI)(OCHO)]+ BArF− (97) | 2.034 | 1.976 | 0.17 | 82 |
26 | [PdII(4-MeBTI)(OC(NHPh)2)]+ BArF− (98) | 2.045 | 1.962 | 0.19 | 82 |
26 | [PdII(4-MeBTI)(SMe2)]+ BArF− (99) | 2.024 | 2.009 | 0.36 | 82 |
26 | [PdII(4-MeBTI)(SeMe2)]+ (100) | 2.020 | 2.010 | 0.36 | 82 |
26 | [PdII(4-MeBTI)(OAc)] (101) | 2.046 | 1.978 | 0.21 | 83 |
26 | [PdII(4-MeBTI)(Cl)] (102) | 2.038 | 1.998 | 0.33 | 83 |
26 | [PdII(4-MeBTI)(PMe3)] B(ArF)4 (103) | 2.018 | 2.036 | 0.44 | 84 |
26 | [PdII(4-MeBTI)(tBuNH2)] B(ArF)4 (104) | 2.030 | 1.988 | 0.34 | 84 |
26 | [PdII(4-MeBTI)(PhNH2)] B(ArF)4 (105) | 2.031 | 1.974 | 0.28 | 84 |
26 | [PdII(4-MeBTI)(Py)] B(ArF)4 (106) | 2.064 | 1.973 | 0.03 | 84 |
26 | [PdII(4-MeBTI)(2,6-Me2py)] B(ArF)4 (107) | 2.055 | 1.969 | 0.01 | 84 |
26 | [PdII(4-MeBTI)(MeCN)] B(ArF)4 (108) | 2.044 | 1.961 | 0.28 | 84 |
26 | [PdII(4-MeBTI)(MeNC)] B(ArF)4 (109) | 2.048 | 2.000 | 0.27 | 84 |
27 | [PdII(3,6-MepMeOi)(OAc)] (110) | 2.078 | 1.957 | 0.27 | 85 |
27 | [PdII(3,6-Me2-BPI)(OAc)] (111) | 2.082 | 1.949 | 0.27 | 41 |
28 | [PdII(iPrBTI)(OAc)] (112) | 2.027 | 1.989 | 0.21 | 85 |
29 | [PdII(6-Me-bpmi)(OAc)] (113) | 2.054 | 1.958 | 0.02 | 41 |
30 | [PdII(4-tBu-BTI)] (114) | 1.986 | 2.076 | 0.03 | 86 |
30 | [PdII(4-tBu-BTI)(Py)] (115) | 2.010 | 2.132 | 0.09 | 86 |
The small number of platinum complexes (Table 7) have similar structural motives to the ones in the previous chapter. Most of them was presented by Wen et al.19,87 They have square-planar geometry around the metal ion with chloride and aromatic coligands. Photophysical properties like luminescence are in the focus of recent research in both biologic (ion sensors) and industrial solutions (photocatalytic hydrogen production).
Ligand | Complex | NPy–Pta (Å) | Nind–Ptb (Å) | Φ | Ref. |
---|---|---|---|---|---|
a The average pyridyl N–Pt distance.b The isoindoline N(H)–Pt distance. | |||||
7 | [PtII(BPI)pyridine](PF6) (116) | 2.058 | 1.964 | 0.15 | 87 |
7 | [PtII(BPI)(C![]() |
2.046 | 1.985 | 0.10 | 19 |
7 | [PtII(BPI)(C![]() |
2.050 | 1.995 | 0.13 | 19 |
7 | [PtII(HL3BPI)(Cl)] (119) | 2.039 | 1.977 | 0.11 | 19 |
18 | [PtIICl(4-Me-10-tBuBPI) (120) | 2.055 | 1.966 | 0.12 | 81 |
Fig. 8 shows a good correlation between tetrahedricity factor (Φ) and endocyclic amine–palladium distances. The increasing central Nind–Pd distance is in agreement with the degree of distortion that is nearly negligible in 107 and the highest in 103. The unique structure of 115 and the N,N,S coordination of 114 likely explain their deviation from the well-observed trend.
The small structural differences of platinum complexes resulted in similar Φ values with 5% deviation (Fig. 8 inset). Distorting factors have a minimal influence on the structures so the trend can be considered zero. This indicates that neither anions nor ligands–coligands can influence the electronic structure of Pt2+.
Fig. 9 shows how tetrahedricity factor (Φ) has been calculated in case of palladium complexes from the opposite planes of a molecule. The method was described on Scheme 3 however the diversity of Pd complexes allows an expressive representation. The angle between planes 1–2 (ϕ1) and planes 3–4 (ϕ2) is small as expected for square planar complexes and increases as distortion grows towards the tetrahedral form. The largest angle (ϕ0) is close to 90° for both SP-4 and T-4 in an ideal case.
Ligand | Complex | NPy–Cua (Å) | Nind–Cub (Å) | τ (Φ) | Ref. |
---|---|---|---|---|---|
a The average pyridyl N–Cu distance.b The isoindoline N(H)–Cu distance. | |||||
7 | [CuII(BPIH)(NCCH3)(OClO3)] (121) | 2.020 | 1.911 | 0.21 | 46 |
7 | [CuII(4-MeBPI)(OAc)] (122) | 2.019 | 1.884 | — | 88 |
7 | [CuII(ind)(pga)(H2O)] (123) | 2.027 | 1.902 | 0.21 | 47 |
7 | [CuII(ind)(mco)] (124) | 2.041 | 1.905 | 0.38 | 89 |
7 | [CuII(NBAII)(OAc)] (125) | 2.003 | 1.895 | — | 90 |
8 | [Cu2II(4′Me-ind)2(MeOH)(μ-OH)]ClO4 MeOH (126) | 2.017 | 1.920, 1.906 | 0.36 | 91 |
8 | [CuII(4′Me-ind)(H2O)2]ClO4 (127) | 1.979 | 1.904 | 0.66 | 92 |
9 | [CuII(bimind)(pga)] 2 DMF (128) | 1.947 | 1.934 | — | 47 |
9 | [CuII(bimind)(ba)] DMF (129) | 1.950 | 1.931 | — | 47 |
9 | [CuII(bimind)OAc] (130) | 1.967 | 1.949 | — | 66 |
10 | [CuII(BTI)(OAc)] (131) | 1.968 | 1.960 | — | 88 |
10 | [CuII(BTI)2] (132) | 2.181 | 2.001, 1.983 | — | 46 |
11 | [CuII(3′Me-ind)2] (133) | 2.265 | 1.952, 1.948 | — | 46 |
13 | [CuII(Mebimind)(Cl)]·0.5CH2Cl2 (134) | 1.956 | 1.945 | 0.44 | 46 |
15 | [CuII(benzBTI)(Cl)] C7H8 (135) | 1.987 | 1.928 | 0.48 | 46 |
31 | [CuII(tetraphenyl-pinBPI)(OAc)] (136) | 2.094 | 1.883 | — | 13 |
32 | [CuII(Me-pentBPI)(OAc)] (137) | 2.038 | 1.912, 1.905 | — | 93 |
33 | [CuII(Bn-pentBPI)(OAc)] (138) | 2.050 | 1.898 | — | 93 |
32 | [CuII(OMe-pentBPI)(Cl)] (139) | 2.029 | 1.913 | 0.36 | 93 |
33 | [CuII(Bn-pentBPIH)(CF3SO3)2] (140) | 1.941 | 1.902 | 0.52 | 93 |
33 | [CuII(Bn-pentBPI)(CF3SO3)] H2O (141) | 1.932 | 1.887 | 0.58 | 93 |
Cu(3′MePyOIND)2 (142) (Fig. 10) belongs to the unsyBII family with a tetrahedral arrangement of ligands around the copper center.23 The complex was found to exhibit catecholase activity by catalyzing the conversion of 3,5-di-tert-butylcatechol to 3,5-di-tert-butylbenzoquinone. Same activity and similar catalytic strength was found for 126 however their coordination sphere is different. Certainly, an incomplete third arm on BII gives the advantage of tunable (hydrophilicity, hydrophobicity, chirality, etc.) 2:
1 complexes while enough space remains for substrate accessibility. Other unsymmetric bidentate isoindoline compounds have been reported to possess antiproliferative activity (143, 144).10
![]() | ||
Fig. 10 UnsyBII ligands and a complex with copper.10,23 |
Stability and redox inactivity have been proved of OC-6 2:
1 complexes (132, 133 and derivatives with ligands: 7, 8, 9, 13, 15). Their 1
:
1 counterparts (121, 134, 135 and derivatives with the same ligands) effectively participated in superoxide scavenging reactions mimicking SOD enzymes.46
The stable d10 electron configuration of Zn2+ has a strong impact on complex formation with BIIs. Conventional methods like using triethylamine base in order to deprotonate the ligand are usually insufficient.69,92,94 The two structures of 147 and 148 (Table 9) have been obtained with tetra-n-butylammonium hydroxide, although some parts in 148 may have remained nondeprotonated.94 151 and 152 have been synthesized in situ from the ligand precursor using the metal salt as catalyst.
Ligand | Complex | NPy–Zna (Å) | Nind–Znb (Å) | τ | Ref. |
---|---|---|---|---|---|
a The average pyridyl N–Zn distance.b The isoindoline N(H)–Zn distance. | |||||
7 | [ZnII(ind)2] (145) | — | — | — | 21 |
7 | [ZnII(NBAII)(OAc)] (146) | — | — | — | 90 |
8 | [ZnII3(4′Me-ind)4] (ClO4)2·5H2O (147) | 2.020 | 1.930 | — | 92 |
8 | [ZnII(4′Me-ind)2] (148) | 2.243 | 2.053, 2038 | — | 94 |
9 | [ZnII(bimind)OAc DMF] (149) | 2.083 | 2.064 | — | 66 |
10 | [ZnII(BTI)(OMe)3] (150) | 2.094 | 2.056 | — | 59 |
20 | [ZnII(phbox)2] (151) | 2.168 | 2.121 | — | 95 |
34 | [ZnII(iPrbox)2] (152) | 2.170 | 2.174 | — | 95 |
The lack of complete deprotonation occurred in 155 (Table 10). The flexibility of the isoindoline system allows the shuttling of the iminium proton between imine N atoms. This intramolecular proton transfer was followed by 1H NMR spectroscopy.94,97 The three Cd structures (153–155) with their preparative circumstances perfectly represent the coordinative capacities of monoanionic aroBIIs.
The last paper to be overviewed is an individual work on a mercury compound by Wicholas' group. The coordination pattern of Hg resembles 153 that is not surprising because similar reaction conditions have been applied. The outer sphere solvent and the inner sphere anion coligand are the dissimilarities [Hg3II(4′Me-ind)4] (NO3)2 4 MeOH.96
As a closure of this section regularities in complex stability have been drafted with complex series of the same ligands. The stability of isoindoline complexes is rarely determined by the K constant. Fortunately, a stability order was found by Irving and Williams for high-spin complexes of divalent metal ions in the middle of the previous century.98 The combination of knowledge in Jahn–Teller effect, crystal field stabilization energy and ionic radius is the key to explain the series. Since it stands for a wide variety of ligands, it was found to be suitable for ligands 7 and 10 as well (Fig. 11). Bond distances of isoindoline Nind–M2+ and NPy–M2+ show the same pattern with the Irving-Williams series.
Ligand | Complex | M–M (Å) | Nind–M (Å) | Ref. |
---|---|---|---|---|
7 | [CuII2(PAP)Cl3(OH)]·1.5H2O (156) | — | 2.000–2.038 | 14 and 100 |
7 | [FeII2(μ-OMe)2(PAP)Cl4]·2MeOH (157) | 3.021 | 2.204 | 101 |
7 | [CoII2(PAP)3(OH)2]Br4·9H2O (158) | 2.795 | 1.880 | 102 |
7 | [NiII2(PAP)3(H2O)2]Br4·6H2O (159) | 3.390 | 2.075, 2.084 | 102 |
35 | [CuII2(PAP-4,6Me)4Cl4] (160) | 3.251 | 2.022 | 103 |
7 | [CuII4(PAP)2(μ2-1,1-N3)2(μ2-1,3-N3)2(μ2-CH3OH)2(N3)4] (161) | 3.207 | 2.021 | 104 |
11 | [CuII4(PAP-3Me)2(μ2-1,1-N3)2(μ2-1,3-N3)2(H2O)2(NO3)2] (162) | 3.163 | 2.009, 1.970 | 104 |
7 | [NiII(PAPH)2(H2O)2](BF4)4 (163) | — | 2.064 | 105 |
7 | [NiII2(PAP)3(H2O)2](NO3)4·5H2O (164) | 3.126 | 2.070, 2.059 | 105 |
7 | [CoIII2(PAP)3(OH)2](NO3)4·xH2O (165) | 2.777 | 1.89, 1.91 | 105 |
7 | [NiII3(PAP)3(OH)2](BF4)3F·xH2O (166) | 2.997–3.026 | 2.048–2.057 | 105 |
13 | [FeIII3(MebimPAP)3O]+ (167) | 3.305–3.33 | 2.047–2.066 | 106 |
14 | [FeIII2(μ-OMe)2(5MeBTI-PAPH2)Cl4] (168) | 3.012 | 2.249 | 68 |
With some perspicuous intention the current major research directions will be discussed in two broader categories such as biomimetics and chemical catalysis.
Two possible mechanisms have been proposed as overall conclusion of PHS mimics with isoindoline complexes. One of them is a metal-based oxidation shown as route 1 in Scheme 4. The rate determining step is the ternary complex formation of metal, ligand and dioxygen that is facilitated by the dioxygen–metal interaction. In the second route a radical formation from H2O2 is assumed initiated by the complexes but substrate transformation happens via standard radical pathway. The first available model experiments were carried out by Prati et al.109 and Barry et al.110 around the 1990's. The recent knowledge is still based on the original 3 × 2e− oxidation process however, it has been extended to other metals and with some radical involvement.
![]() | ||
Scheme 4 The proposed mechanism of phenoxazinone synthase model reactions.67 |
Recently, as a major improvement an expressive correlation was discovered (Fig. 12) how fine-tuning of the precursor complex could increase reaction rates during AXP synthesis. Modifications on BII strongly influence the oxidation potential of iron complexes that has a straight effect on the reaction rates.
![]() | ||
Fig. 12 Isoindoline dependent oxidation potentials in iron complexes vs. reaction rates in phenoxazinone synthase mimics.67 |
According to kinetic data a ternary complex formation is plausible among superoxoiron(III), the corresponding ligand and dioxygen (route 1, step 2 in Scheme 4) in the rate determining step.67
The catalase activity of manganese complexes has been determined by gas-volumetric method measuring dioxygen evolution from H2O2 in specific time intervals. Results are shown in Table 13 compared to the original enzyme efficiency from various sources. The kcat/KM values of models are significantly lower than catalase, but they are in the same range with other model compounds.117 The most active compound up to this date is a Mn4+ complex with a SALPN type ligand that is still far behind the native enzyme.118 The hypothetical cycle of mechanism is explained by Mn3+–Mn4+ (mononuclear) or Mn2+/Mn2+–Mn3+/Mn3+ (dinuclear) transitions.118,119
Considering future directions of developments some useful observations have been made how coligands and other structural changes are able to influence activity. Inactive 1:
1 or slightly active 1
:
2 (metal–ligand ratio) manganese complexes could be enhanced by active molecules such as TEMPO (2,2,6,6-tetramethyl-piperidin-1-yl-oxyl)120 or different N-donor heterocycles (pyridine, imidazole)17 (L). Electron-donating moieties on both coligands (changing imidazole to 1-methylimidazole)17 and R-groups of 3 (more effective 9 than 7)50 have increasing effect on activity. On the basis of reaction kinetic data a mechanism has been proposed where Mn(II) is transformed into a Mn(IV)
O species that is reduced by H2O2 to its initial form at the end of the catalytic cycle (Scheme 6).17
![]() | ||
Scheme 6 The proposed mechanism of catalase model reactions.17 |
McCord–Fridovich assay125 is a well-known spectrophotometric protocol that is used to determine superoxide dismutase activity. All complexes presented in Table 14 have been tested in a reaction of nitroblue tetrazolium (NBT) or cytochrome c reduction. The superoxide radical anion (O2˙−) was generated in situ by the xanthine/xanthine oxidase reaction, and its decomposition was monitored spectrophotometrically by following the formation of diformazan from NBT at 560 nm, or alternatively, by monitoring the reduction of cytochrome c at 550 nm. The products of the disproportionation are dioxygen and hydrogen peroxide.
Ligand | Complex | Epca (Eox/red) | IC50 (10−6 M) | kMcF (kcat) (10−6 M−1 s−1) | Ref. |
---|---|---|---|---|---|
a Redox potential values are expressed in Volt. | |||||
— | CuZnSOD | — | — | 2000 | 126 |
9 | [CuII(bimindH)(NCCH3)(OClO3)]ClO4 (171) | −0.835 | 6.62 | 1.98 | 46 |
9 | [CuII(bimindH)Cl2] (172) | −0.835 | 10.99 | 1.19 | 46 |
13 | [CuII(Mebimind)Cl]·0.5CH2Cl2 (134) | −0.860 | 27.37 | 0.48 | 46 |
10 | [CuII(BTIH)(NCCH3)(OClO3)]ClO4 (173) | −0.826 | 4.69 | 2.80 | 46 |
10 | [CuII(BTI)Cl] (174) | −0.826 | 7.19 | 1.83 | 46 |
7 | [CuII(BPIH)(NCCH3)(OClO3)]ClO4 (121) | −0.724 | 0.95 | 13.82 | 46 |
7 | [CuII(indH)Cl2] (175) | −0.724 | 2.06 | 6.37 | 46 |
15 | [CuII(benzBTI)Cl] (135) | −0.650 | 1.21 | 10.85 | 46 |
10 | [FeII(BTI)2] (176) | (0.555) | — | (5.00) | 22 |
10 | [FeII–III(BTI)2] MeCN (55) | — | — | (5.28) | 22 |
7 | [FeII(ind)2] (47) | (0.319) | — | (3.83) | 22 |
8 | [FeII(4′Me-ind)2] (177) | (0.265) | — | (3.62) | 22 |
9 | [FeII(bimind)2] (52) | (0.205) | — | (3.06) | 22 |
13 | [FeII(Mebimind)2] (178) | (0.123) | — | (2.50) | 22 |
9 | [MnII(bimind)2] (179) | (−0.021) | 3.10 | (3.28) | 15 |
13 | [MnII(Mebimind)2] (180) | (0.005) | 6.36 | (1.60) | 15 |
10 | [MnII(BTI)2] (45) | (0.082) | 12.10 | (0.83) | 15 |
7 | [MnII(ind)2] (181) | (0.015) | 8.74 | (1.26) | 15 |
11 | [MnII(3′Me-ind)2] (182) | (−0.098) | 11.10 | (0.99) | 15 |
8 | [MnII(4′Me-ind)2] (183) | (−0.082) | 9.44 | (1.08) | 15 |
7 | [CoII(ind)2] (184) | −0.483 | 4.29 | 2.35 | 53 |
11 | [CoII(3′Me-ind)2] (185) | −0.512 | 4.82 | 2.10 | 53 |
8 | [CoII(4′Me-ind)2] (186) | −0.583 | 15.90 | 0.64 | 53 |
SOD models are categorized by inhibition constant IC50 values (nM to μM range, the concentration where inhibition of NBT or cyt c reduction is 50%). Unlike with other enzyme models the primary aim here is the invention of an effective SOD substitute. Kinetic measurements focus on the achievement of lower IC50 or higher kMcF instead of the understanding of the mechanism. Rather new findings with cobalamins may point to the right direction namely their efficacy approach SOD.127
Overviewing the results of SOD mimicking activity it is clear that the transition metal complexes of BIIs are relevant artificial models of SOD function. Redox potentials, coordination geometry and ligands have the key roles in the process. Octahedral complexes with irreversible redox behavior have a little contribution to superoxide destruction.46,53 Scenic correlation was found between reduction peak potential (Epc) and inhibition constant at copper complexes.46 Iron activities correlate better with redox potentials (Eox/red)22 and manganese complexes are influenced by counter-anions and non-coordinating ligand parts.15 Certainly, the redox properties of metals show the major influence that was also proved by comparing superoxide scavenging reaction of Mn, Fe, Co and Ni complexes of the identical ligand (7). The highest activity of Fe was observed at that time (Fig. 13).53 Although, the series is not complete (missing data for homoleptic complex of Cu) it is clear that iron and copper have promising values at least among the SOD models (see also Fig. 14).
![]() | ||
Fig. 13 Rate constants for the superoxide scavenging reaction of 1![]() ![]() |
Fig. 14 shows the kMcF rate constants plotted against the E°′1/2 for all complexes that have been investigated as SOD models. A clear trend is observable in case of Cu and Fe while Mn and Co act differently. Redox potential values can be set where SOD activity is optimal. These are around the CuZnSOD redox couple (∼0.06 V vs. Ag/AgCl in 1 M KCl at pH 7.5 in HEPES buffer)128 and above the O2/O2˙− potential (−0.40 V vs. SCE – saturated calomel electrode).129–131
The N-donor coordination sphere built by BIIs is an accurate model environment of CuZnSOD's copper, where three histidine ligands surround that part of the active center. As it was mentioned earlier cobalamins127 can approach SOD activity implying to further ligand effects that have to be taken into consideration for future modelling studies.
Model experiments are usually carried out with 3,5-di-tert-butylcatechol to avoid dimerization. Products are 3,5-di-tert-butyl-1,2-benzoquinone (3,5-DTBQ) and H2O2. As it is expected from model systems the activities are several magnitudes lower compared to the enzyme. The slow transformation is fortunate for reaction kinetic observations. The kcat/KM and TON values of Table 15 give the general information on substrate conversion during time unit. Turnover rate differences show that binuclear PAP complex have better mimicking properties at the same time mononuclear BIIs are more suitable for the investigations of influential factors. The only binuclear copper complex (126) is the most accurate catecholase structural model. The Cu–Cu distance is 3.274 Å that makes catechol coordination possible.
Ligand | Complex | TON (h−1) | kcat/KM (M−1 s−1) | Ref. |
---|---|---|---|---|
— | Catechol oxidase from I. batatas | — | 8250 | 133 |
12 | [Mn2(6′Me2PAP)2Cl4] (187) | 167.9 | 7.25 | 99 |
12 | [Mn(6′Me2indH)(H2O)2(CH3CN)](ClO4)2 (188) | 48.8 | 1.07 | 16 |
7NMe | [CuI(Me2ind)I2] (189) | 65.0 | 67.40 | 33 |
8 | [Cu2II(4′Me-ind)2(MeOH)(μ-OH)]ClO4 MeOH (126) | — | 0.749 | 91 |
11 | [Cu(3′MePyOIND)2] (142) | 21.5 | 23.37 | 23 |
According to the enzymatic met (native) and oxy form the ideal distance has to be between 2.9 Å and 3.8 Å. The enzyme like behavior of models is attributed to the binuclear cores (with the appropriate distances) and to the presence of enzyme-like N-donor environment (Fig. 15).
![]() | ||
Fig. 15 Catechol oxidase active site (left), the coordination environment of [Cu2II(4′Me-ind)2(MeOH)(μ-OH)]ClO4 MeOH (126) (middle)91 and the proposed structure of [Mn2(6′Me2PAP)2Cl4] (187) (right).99 |
Observations from the kinetic experiments imply to a biocatalytic circle where a ternary complex of dioxygen, substrate and catalyst is involved, furthermore bases and electron donating potential of the substrate have an effect on the mechanism. Basicity is considerable from both catalyst and external side as the prior deprotonation of the catechol is required for the activation of the “dezoxy form” of the catalyst.16,23
![]() | ||
Scheme 8 The proposed mechanism of catechol 1,2-dioxygenase model reactions (top) and the effect of fine-tuning on the k reaction rate (bottom).65 |
The reaction mechanism of catechol dioxygenase mimics hypothesize certain active iron–oxygen species that can materialize primarily from iron(III). In the most recent works of our group we had special attention on stable 1,2-peroxodiiron(III) intermediates.34,68 These species have been identified in several industrially promising enzymes (e.g. soluble methane monooxygenase, ribonucleotide reductase or desaturases) by Raman and EXAFS spectroscopy,134,135 furthermore by characteristic UV-Vis vibrations at 600–750 nm closer to the near-infrared region. Oxidation reactions of thioanisoles and benzyl alcohols have been investigated in the presence of H2O2 using mild reaction conditions aiming for selectively oxidized products. The work presented in these papers contains proofs of a metal-based oxidant with the clear exclusion of hydroxyl radicals. Reactions with para-substituted reagents showed changes in oxidation rate as a function of the nature of the para substituent. The negative Hammett ρ values were indicative of an electrophilic oxidant, furthermore KIE (kinetic isotope effect) supported the theory about the lack of radical involvement. Results are a part of a continuous research on the role of metal-based oxidants in non-heme diiron enzymes.
Palladacycles are known with unsyBIIs as well.136,137 Caryl–H bond activation was observed in this case.
Catalytic hydrogenation with palladium has been carried out with more efficacy on aliphatic olefins.18 It is notable that the catalyst could be reused in several cycles.
Multiple papers from L. H. Gade's group describe CC bond activation. The peroxylation of cyclohexene to tert-butylperoxy-3-cyclohexene has been tested with copper, cobalt and iron complexes.61,88 Selectivity towards the desired product was reported to be outstanding. Using highly reactive oxaziridine type oxidizing agents even epoxidation products have been detected from a great number of olefins by iridium-aroBIIs (Scheme 9).78
Cobalt(III)–alkylperoxy complexes can be activated by high energy light irradiation. Oxidation products of alkanes or alkenes are alkylperoxides or ketones/alcohols depending on the presence of inert atmosphere.138
Quantum yields dramatically change when luminescence of complexes is investigated. It dropped below 5% when platinum or rare earth metals have been measured.19,32,46
Specific optical properties have a wide range of applications in modern electronic devices. Crystallographic direction dependent refractive index (birefringence) has been determined for a few aroBII ligands. These values describe the 3D structure of a crystalline material giving information on anisotropic polarizability or supramolecular orientation, etc.139
UnsyBIIs have already been discussed (see above) in their active palladacycle form and as bidentate N-donor ligands. Recently, antiproliferative activity has been tested too on five human tumor cell lines. The results serve as guideline for future development of bioactive isoindoline derivatives.10
This journal is © The Royal Society of Chemistry 2015 |