Next Article in Journal
Polymeric Membranes for Oil-Water Separation: A Review
Previous Article in Journal
Cytotoxicity, Colour Stability and Dimensional Accuracy of 3D Printing Resin with Three Different Photoinitiators
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Heterogeneous Dendrimer-Based Catalysts

1
Department of Petroleum Chemistry and Organic Catalysis, Moscow State University, 119991 Moscow, Russia
2
Institute of Petrochemical Synthesis RAS, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(5), 981; https://doi.org/10.3390/polym14050981
Submission received: 28 January 2022 / Revised: 22 February 2022 / Accepted: 23 February 2022 / Published: 28 February 2022
(This article belongs to the Topic Multiple Application for Novel and Advanced Materials)

Abstract

:
The present review compiles the advances in the dendritic catalysis within the last two decades, in particular concerning heterogeneous dendrimer-based catalysts and their and application in various processes, such as hydrogenation, oxidation, cross-coupling reactions, etc. There are considered three main approaches to the synthesis of immobilized heterogeneous dendrimer-based catalysts: (1) impregnation/adsorption on silica or carbon carriers; (2) dendrimer covalent grafting to various supports (silica, polystyrene, carbon nanotubes, porous aromatic frameworks, etc.), which may be performed in a divergent (as a gradual dendron growth on the support) or convergent way (as a grafting of whole dendrimer to the support); and (3) dendrimer cross-linking, using transition metal ions (resulting in coordination polymer networks) or bifunctional organic linkers, whose size, polarity, and rigidity define the properties of the resulted material. Additionally, magnetically separable dendritic catalysts, which can be synthesized using the three above-mentioned approaches, are also considered. Dendritic catalysts, synthesized in such ways, can be stored as powders and be easily separated from the reaction medium by filtration/centrifugation as traditional heterogeneous catalysts, maintaining efficiency as for homogeneous dendritic catalysts.

Graphical Abstract

1. Introduction

Various catalytic processes, such as hydrogenation, oxidation, polymerization, etc., are of great importance in the modern petrochemistry and pharmaceutical industry [1,2,3,4,5,6]. The last tendencies, concerning the advanced science and technology in catalysis, consist in the transition to nanoscale and nanoreactor supramolecular systems, providing higher reaction rates and selectivity due to the metal nanoparticle high specific surface area, size, geometry, and electron effects as well as of specific microenvironment metal complexes or nanoparticles [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25]. To prevent metal nanoparticle sintering and agglomeration, and, as a consequence, catalyst deactivation during the process, various organic ligands, such as ionic liquids [26,27,28,29,30,31,32,33,34,35], simple amines, phosphines, and thiols [16,26,27,28,29,30,31,32], cyclodextrins [24,26,35,36,37,38,39,40,41,42,43], calixarenes [26,44,45,46], and polymers were suggested [24,26,27,28,29,30,31,32,35,37,38,39,43,47,48,49,50,51,52,53,54,55,56]. Worthy to distinguish among the latter are dendrimers, three-dimensional globular macromolecules with a branched regular structure [10,57,58,59,60,61,62,63,64].
Dendrimers were independently synthesized for the first time by Vögtle [65,66], Denkewalter [67], Tomalia [68], Fréchet [69] and Newkome [70] in the 1980s. Since that time, due to their unique structure and relative facility for modification, dendrimers have found a wide application in nanomedicine [25,61,62,63,71,72,73,74,75,76,77,78], light and electron-sensing devices [25,63,71,72,76,79,80,81,82,83], material science [10,24,25,61,62,63,64,72,76,77,78,79,80,82,83,84,85,86], and, especially, in nanocatalysis [10,24,25,26,28,60,61,62,63,64,72,79,81,87,88,89,90,91,92,93,94,95]. Hence there were developed materials, based on poly(amido amine) (PAMAM) (Tomalia type) [68,96], poly(propylene imine) (Vögtle type) [66,81,97], poly(ether imine) (PETIM) [98,99,100,101,102], poly(ester amide) (Newkome type) [70,103], poly(benzyl/aryl ether) (Fréchet type) [68,104], poly(aliphatic ester) (Newkome type) [105], poly(carbosilane) [106], triazol or allyl-ended carbosilane arene-cored (Astruc type) [107,108], phenylazomethine (DPA) (Yamamoto type) [80,109], arylphosphite phosphazene cored (Caminade/Majoral-type) dendrimers, etc. (Figure 1) [110,111,112,113].
In the dendrimer structure, there is a core, branches, and periphery (terminal, end, or surface) groups (Figure 2) [10,26,63,64,68,71,79,80,85]. Herein PAMAM dendrimers of higher generations (G) are comparable in the size (up to 10 nm) with protein molecules [75,84,87,114].
The main advantages for dendrimer application in catalysis are: (1) the possibility to sorb a well-defined quantity of metal through the coordination by donor node and end groups [10,26,60,62,70,81,82,83,91,93,115]; (2) the control of substrate selectivity via choice of the dendrimer appropriate generation [25,26,60,92,93,116,117,118] and/or by terminal group modification, specifying the dendrimer solubility in the reaction medium and substrate affinity [25,60,62,73,80,90,92,93,116,117,118]; (3) the recyclability of dendrimer-based catalysts via fractional precipitation from solution or separation to another phase by the addition of appropriate, so-called “bad” solvents [60,89,93,108,115,119,120,121,122,123,124,125,126,127,128,129,130], or cooling the reaction mixture below upper critical solution temperature (UCST) or heating above low critical solution temperature (LCST) [63,71,77,89,90,94,115,131,132,133,134,135]. The active metallocenters in dendrimers can be located in a core or at the focal point (Figure 3a) [10,60,62,64,79,84,89,90,92,93,113,119,120,121,129,131,136,137,138], at nodes (Figure 3b) [10,60,64,80,93,113,132,136,139,140], end functional groups (Figure 3c) [60,63,79,89,90,92,93,94,101,102,108,113,136,140,141,142,143,144,145,146,147,148,149,150], or in a cavity between dendrimer branches (Figure 3d) [10,60,62,64,90,93,94,115,116,117,118,125,126,136,139,150].
In the end of 1990s and 2000s, Crooks and co-workers [10,25,26,28,60,62,64,91,92,115,117,151,152,153,154], on the one hand, and Kaneda and Mizugaki [10,26,28,62,64,91,92,116,155,156,157], on the other hand, suggested an approach for the synthesis of Cu, Pt, Pd, Rh, Ag, and Au nanoparticles (NP) of 1–3 nm in diameter, encapsulated inside the cavities of PAMAM or PPI dendrimers. This approach includes two steps (Scheme 1): (1) metal ion complexation with the dendrimer functional groups (both nodes and periphery); and (2) reduction by NaBH4 or KBH4 (sodium or potassium borohydride), resulting in nanoparticle or cluster formation. It can be applied for synthesis of both monometallic and bi- or polymetallic dendrimer-encapsulated nanoparticles [10,25,62,64,91,92,95,115,158].
In the last case, one may distinguish three different sub-approaches (Scheme 2). The first one is co-complexation of two or more different metal ions simultaneously, resulting in alloy or core-shell nanoparticles depending on the reduction rate of each metal (Scheme 2) [10,62,64,88,91,92,95,115,131,154,158,159,160,161,162,163,164,165,166]. The second one is the subsequent complexation, including two repeated procedures of complexation followed by reduction for each metal and, as a rule, resulting in core-shell nanoparticles (Scheme 2) [10,62,64,88,91,95,115,154,158,167,168]. The third one is the so-called galvanic displacement (Scheme 2), suggesting the use of oxidizable metal (e.g., Cu0) as a reducing agent for more noble metals with higher oxidation-reduction potential (e.g., Pt, Au), resulting in core/shell or alloy-type bimetallic nanoparticles [88,91,95,115,158,169,170].
The driving force, providing the metal ion adsorption, uniform distribution, and well dispersion in the interior of the dendrimer molecule, further resulting in the formation of metal nanoparticles and clusters with narrow size distribution, was here for the coordination and complex formation between metal ions and dendrimer amino groups [60,88,115,116,117,151,155,156]. Moreover, dendrimer-encapsulated nanoparticles and clusters may be formed from dendritic metal complexes in situ under hydrogenation conditions [140,148,171] or via partial reduction by NH2 end groups of PPI or PAMAM dendrimers [128,148,171,172,173,174,175,176,177]. As a rule, due to the steric hindrances, the mean particle size decreased with the increase in dendrimer generation [91,116,155].
The similar approach, including metal ion complexation followed by reduction with NaBH4, was applied by Yamamoto’s group for the synthesis of ultra-small mono-, bi-, and polymetallic nanoparticles and clusters (of 0.5–1.5 nm in diameter), encapsulated into the cavities of poly(phenylazomethine) dendrimers [10,64,162,178,179,180,181,182,183,184]. It is worth noting that, in contrast to globular PPI, PAMAM, and other related poly(alkyl amine) dendrimers, the complexation of rigid poly(phenylazomethine) dendrimers (DPA) with metal ions proceeds radially, layer by layer, from core to periphery (Scheme 3) [10,64,80,109,166,178,179,180,181,182,183,184,185]. This is due to the rigid, conjugated, electron transfer structure of poly(phenylazomethine) dendrimers, resulting in higher electron density and, as a consequence, in higher proton association constants for core imine groups, as compared with periphery [10,80,109,162,183,184,185].
With regard to bi- and polymetallic complexes, it implies the preferential complexation of inner imine groups by metal ions with stronger association constants, whereas less affine metal ions coordinate outer imine sites [10,109,166,183,184,185]. Reduction of bi- and polymetallic DPA complexes with sodium borohydride afforded to formation of alloy-type nanoparticles and clusters [166,183,184,185].
Astruc et al. developed hybrid aryl-cored poly(carbosilane/triazol/aryl ether) dendrimers, terminated with allyl, ferrocenyl, triazol, sodium sulfonate, or tri(ethylene glycol) moieties, and used them for the synthesis of Pd, Cu, and Au nanoparticles with sodium borohydride, MeOH, or sodium ascorbate as reducing agents [90,107,108,150,186,187,188,189,190,191]. Herein, depending on the dendrimer size and generation, as well as on the metal to dendrimer ratio, small dendrimer-encapsulated nanoparticles (DENs) of 1–2 nm can be obtained for Pd with the narrower size distribution or dendrimer-stabilized nanoparticles (DSNs) with the sizes of 1–4 nm for Pd, 3–12 nm for Au, and 4–20 nm for Ag (Figure 4) [10,25,62,64,90,150,186,187,188,191]. The similar is also true for PPI, PAMAM, and other related poly(alkyl amine) dendrimers [60,115,116,156,168,192,193].
Dendrimer-stabilized core nanoparticles (or nanoparticle-cored dendrimers, NCD) were mostly synthesized in the presence of Fréchet type poly(aryl ether) or poly(aryl ester) dendrons with the donor phosphine, thiol, hydroxyl, or amino groups at the focal point, resulting in so-called monolayer protected cores (MPC) (Figure 5) [10,62,84,89,119,120,121,129,137,194,195,196,197,198,199,200]. The synthesis was usually carried out in the organic solvent (toluene or THF) as a metal source and reducing agent, respectively; the combinations of easily decomposed organic metal salt (e.g., Pd(OAc)2) or complex (e.g., [Ru(Cumene)Cl2]2 or [Ru(COD)(COT)]) and molecular hydrogen [119,120,129], or chloride complex salt or acid (e.g., K2PdCl4, HAuCl4 or H2PtCl6), often additionally stabilized by amphiphilic ionic liquid (e.g., [N(C8H17)4]Br), and sodium borohydride [121,137,194,195,196,197,198,199,200], were used. Another approach implied the modification of nanoparticles, already stabilized by bifunctional ligands (e.g., Au NPs, stabilized by 11-mercaptoundecanoic acid), with hydroxyl poly(aryl ether) dendrons of various generations (Scheme 4), hence providing the nanoparticle uniform and narrow in size distribution [201].
Herein the mean particle size was strongly dependent on the nature of the donor group in the dendron focal point, reaching 1–6 nm for Au and Pd nanoparticles, coated with more affine thiol or phosphine dendrons [121,129,194,195,196,197,198], thus revealing the wider particle size distribution in comparison with PAMAM or PPI dendrimer-encapsulated nanoparticles [91,116,151,156], and 2–25 nm for Au-cored dendrimers with hydroxyl stabilizer at the focal point [199], and are thus comparable with monometallic Pd and Au nanoparticles, stabilized by PAMAM, PPI [168,192], or poly(carbosilane/triazol/aryl ether) dendrimers [188,191]. Pd nanoparticle core size decreased with the increase in the dendron generation [129], whereas Au and Ag nanoparticle core size increased [195,196,197,198,199,202], with the distribution, depending not only on the dendron generation but also on the reduction conditions [199].
Among the other nanoparticle-cored dendrimers, the synthesis of dendrimer core monometallic Pt and bimetallic core/shell Pt[Au] nanoparticles was also reported, stabilized by poly(aryl ether) dendrons with tris(acetic acid) ammonium moieties at focal points (Figure 6) [62,200,203]; of Au nanoparticles, stabilized by thiol poly(L-lysine) dendrons (Figure 7) [83,204]; of core/shell CdSe/ZnS nanoparticles, stabilized by thiol-focal allyl-terminated poly(ethylene imine)/poly(amido amine) dendrons, which further were linked together by metathesis reaction (Figure 8) [84,205,206]; of monometallic Pd and bimetallic PtPd alloyed nanoparticles, stabilized by PAMAM dendronized 4-(aminomethyl pyridine), etc. [131,207]. In the last case, both organic metal sources ([PdCl(C3H5)]2 and PtCl2(PhCN)2 in CH2Cl2) and reducing agent (LiB(C2H5)3H) were used, resulting in mono- or bimetallic nanoparticles with strongly reproducible mean sizes for the same dendron generation and periphery and decreasing with the increase in the dendron generation and the length of terminal moiety (Scheme 5) [131,207].
Thus, synthesized metal complexes and nanoparticle dendrimer-based catalysts, both monometallic and bi/polymetallic, were successfully applied in various processes, such as hydrogenation of alkenes, phenols, aromatic and heterocyclic compounds, semihydrogenation of alkynes, dienes and polyenes, hydrogenation and hydroamination of carbonyl compounds, nitroarene reduction, oxygen reduction, hydroformylation, alkyne-azide cycloaddition, allylic addition, asymmetric synthesis and catalysis, Wacker oxidation, C–H, O–H, and sulfide oxidation and epoxidation, olefin oligomerization and polymerization, metathesis, oxidative coupling, and various cross-coupling reactions since the second half of the 1990s until the end of the 2010s (see Table 1) [10,24,25,26,60,62,64,72,79,84,88,89,90,91,92,93,94,113,114,115,136,145]. Herein reaction turnover numbers (TON) reached 2,700,000, while turnover frequencies (TOF) exceeding 28,000 h−1 [186,188]. Moreover, bi- and polymetallic catalysts, both metal complex [109] and nanoparticle [131,159,160,163,164,166,170,183,203], exhibited the enhanced efficiency due to the metal synergism and changed electron properties and geometry of nanoparticles, clusters, and complexes [10,62,109,115,164,183,203,208,209,210,211,212,213].
Depending simultaneously on the reaction type, dendrimer nature, and catalytic centers location, as well as on the overall weight of the stabilization or steric factors from the dendrimer ligand, the so called positive dendritic effect can be observed, appearing in the enhanced catalyst performance with the increase in the dendrimer generation [79,80,81,82,83,84,85,86,87,88,89,90,92,94,113,115,120,121,129,131,132,136,141,145,146,149,190,193,195,207,215,216,218,221,227,242,247,248,252,255,258,261,262,263,271,272,273,274,279,280,281,282,283] or the negative dendritic effect, respectively, and the drop in the catalytic activity with the increase in the dendrimer generation or substrate size [60,90,91,94,101,116,117,122,124,126,130,132,136,142,143,145,147,148,150,187,193,200,214,227,234,271,281,282,283,284].
Moreover, the dendrimer nature was found to have an essential influence on the selectivity in the hydrogenation of alkynes and dienes in the presence of Pd-containing catalysts. When Pd nanoparticle-cored Fréchet type poly(aryl ether) dendrimers were used, both covalently attached to the Pd surface [137] or stabilizing via phosphine focal point [129], C≡C triple bonds or conjugated C=C double bonds underwent exhaustive hydrogenation. Vice versa, alkynes and dienes could be selectively converted to corresponding monoenes in the presence of Pd catalysts (both mono- and bimetallic), based on triazol [150], PAMAM [131,148,159,160,207,241] and, especially, on PPI dendrimers [116,128,140,156,171,193,241], with the total selectivity on alkenes increased with the increase in the dendrimer generation [116,148,193]. It was due to the poisoning effects of pyridyl moieties at the dendron focal point and/or dendrimer node donor tertiary amino group, occupying a part of adsorption sites and thus enriching the nanoparticle surface with electrons, facilitating the de-adsorption of alkene formed and interfering with its possible further re-adsorption [4,128,131,171,177,193,207,285,286,287]. Herein, the product distribution in semi-hydrogenation of alkynes and, especially, dienes, was strongly affected by metal composition of the catalyst active phase; the dendrimer nature, influencing the metal surface electron properties, as well as substrate and intermediate adsorption features; and, finally, the substrate structure (including the accessibility of C=C double bonds) and electronic properties (the presence of electron donor (EDG) or electron withdrawing groups (EWG)) [128,131,150,171,193,207,288,289].
However, in spite of their remarkably high activity, selectivity, and recyclability [121,130,131,132,190,234,269], homogeneous dendrimer-based catalysts can be stored and used mostly as colloidal solutions, undergoing gradual deactivation due to metal leaching and precipitation from the dendrimer stabilizing micelles [118,126,164,186,187,278] or, otherwise, should be prepared in situ [130]. As a consequence, since the end of the 1990s, various approaches for heterogenization of dendrimer-based catalysts have been suggested. These heterogenized dendritic catalysts are supposed to exist as powders, easily precipitate, and isolate from the reaction medium using filtration or centrifugation—similar to conventional heterogeneous catalysts. In some cases, homogeneous micellar dendritic catalysts, initially obtained as solids, may be used as heterogeneous in the “bad” solvent medium [121]. Immobilization of dendritic catalysts, using various heterogeneous carriers, such as silica, polystyrene, etc., allows to significantly enhance their stability and recyclability.
Presently, three main approaches to the synthesis of immobilized heterogeneous dendrimer-based catalysts have been developed and are further considered in the present review.
The first is the adsorption of metal-containing dendrimers or dendrons on the surface of amorphous silica, carbon, or inside channels in channels of mesoporous materials via hydrophobic, electrostatic, or π–π interactions [86]. This approach is considered in Section 2.
The second approach supposes the covalent grafting of dendrimers or dendrons on polymers, amorphous silica, carbon nanotubes, or inner walls of mesoporous materials [86]. It is considered in Section 3.
The third approach, including the dendrimer cross-linking, using various transition metal ions, as well as bi- or trifunctional agents, polymers, etc., [241], is discussed in Section 4.
Additionally, magnetically separable dendritic catalysts [62], which may be designed using all three approaches mentioned above, are considered in Section 5.

2. Non-Covalent Immobilization of Dendritic Catalyst

One of the simplest methods for the immobilization of homogeneous dendritic catalysts, developed by Somorjai [10,64,290,291] and Yamamoto [10,64,178], implies the deposition of colloidal DEN solutions (mono- or bi/polymetallic) onto the various heterogeneous carriers, such as silica (both amorphous and mesoporous), zirconia, and carbon materials, by the wetness impregnation method (Scheme 6). It is worth noting that the said approaches suppose the remaining of dendrimers inside the SBA-15 channels or on the carbon surface using electrostatic, hydrogen bonding, hydrophobic, or π–π interactions, hence preventing nanoparticle agglomeration sintering [178,292,293]—in contrast to previous works, where dendrimers were used just as templates for nanoparticle and pore formation and then removed under calcination [26,169,290,291,294,295,296,297,298,299,300,301]. Thus, well dispersed and tiny nanoparticles of 1.0–2.5 nm in diameter inside the SBA-15 channels [292,293] or subnanoclusters with the sizes of 0.5–1.5 nm, deposited onto mesoporous carbon [178,180,183,184], were obtained.

2.1. Impregnation/Adsorption on Silica-Based Carriers

Using the above-mentioned approach, Somorjai and co-workers synthesized highly active Pt and Rh nanocatalysts for ethylene (at room temperature and 10 atm. of H2) and pyrrole hydrogenation, including ring-opening reaction (at 60–150 °C and 50 atm. of H2) with turnover frequencies (TOF) up to 4.3 s−1 (~15,500 h−1) (Figure 9) [292,302]. Pd, Pt, and Rh nanoparticles, encapsulated into PAMAM dendrimers and supported on SBA-15, appeared highly active in the reversible dehydrogenation/hydrogenation of N-heterocycles, such as (tetrahydro)quinolines and (dihydro) indoles at 130/60 °C, resulting in the quantitative yields within 24 h [303]. Furthermore, PAMAM dendrimer-encapsulated Pt and Rh nanoparticles, supported on SBA-15, were successfully applied for liquid (at 20–80 °C) and gas-phase (at 200−225 °C) hydrogenative ring-opening and isomerization of methylcyclopentane (MCP) and alkyl/aryl cyclopropane (Figure 9) [304,305], whereas Pd nanoparticles, synthesized in the same manner, favored MCP dehydrogenation and ring enlargement to benzene at 250–275 °C, giving quantitative conversions and TOF values up to 335 h−1 [304], as well as effectively catalyzed electrophilic arylation of indole and one-pot tandem dehydrogenation/arylation of indoline at 60–130 °C (Figure 9) [303]. Herein, increases in the mean particle size and dendrimer generation lowered the reaction activation energy, thus favoring the enhanced catalyst performance [305].
Supported on SBA-15, Pt, Pd, and Au nanoparticles encapsulated in PAMAM dendrimers, in the presence of PhICl2 as an oxidative catalyst activator, exhibited high activity, just slightly inferior to that for homogeneous catalysts PtCl2 or AuCl3, in the electrophilic cycloaddition of ortho-alkynyl arenes, phenols, anilines, benzyl amines, or benzoate esters to C≡C triple bond, thus forming 5- or 6-membered heterocycles with the yields of 45–98%, depending on the substrate structure (20–100 °C, 15 h, toluene, Ar) (Figure 9) [306,307,308,309]. It should be noted that Pt@PAMAM-G4/SBA-15 appeared to be a much more stable catalyst in comparison with its analogue, stabilized by poly(vinyl pirrolidone) (PVP), and maintained its efficacy at reuse [306].
Moreover, the catalyst Au@PAMAM-G4/SBA-15 revealed the best performance and well stability in the cyclopropane synthesis via alkene-alkyne cycloaddition (20–70 °C, 12 h, toluene) (Figure 9) [310] and Hayashi-Ito aldol cycloaddition between aryl aldehyde and methyl isocyanoacetate (20 °C, PhICl2, 18 h, toluene), thus surpassing conventional homogeneous catalysts, based on Au carbene complexes (Figure 9) [311]. In the last case it was additionally established that larger pore size of the heterogeneous carrier favored the preferential trans-product formation, whereas smaller pore size constrained cis-product formation respectively, with a common decrease in the reaction rate for +I and −M substituents in the aryl aldehyde molecule, thus retarding electrophilic attack of the aldehyde group to enolate deprotonated forms of isocyanoacetate (Figure 10) [311].
It is worthy to note that heterogenized catalysts, based on PAMAM dendrimer-encapsulated nanoparticles loaded to mesoporous SBA-15 silica, can be easily adopted not only for batch, but also to continuous-flow conditions, exhibiting well time-on-stream and recycling properties [307,309,310].
The similar approach was suggested for the synthesis of PAMAM dendrimer-encapsulated Ir nanoparticles, immobilized on amorphous SiO2, and further applied for the hydrogenation of 2-nitrobenzaldehyde under the mild conditions (30 °C, 1 atm. of H2) [227]. Herein, the conversion of 98% was achieved within 1 h, and only the NO2 group in the substrate structure was subjected to hydrogenation. The increase in mean particle size from 0.9 to 1.5 nm resulted in the increased portion from 17% to 42% of anthranil, the product of 2-nitrobenzaldehyde semihydrogenation followed by a heterocyclization-dehydration reaction between the intermediately formed NHOH group and the C(=)OH group (Scheme 7), whcih was attributed to more efficient dissociative adsorption of H2 on smaller Ir clusters [227].
Meijboom et al. proposed additional post-treatment of PAMAM dendrimer-encapsulated Ru nanoparticles, immobilized on amorphous silica, with ionic liquid solution; thus, the so-called solid catalysts with an ionic liquid layer (SCILL) were obtained [312,313]. The mean particle size was found as 2.1–2.4 nm, decreasing with the increase in the dendrimer generation from G4 to G6. The catalysts synthesized were studied in the hydrogenation of citral (90–130 °C, 10–30 atm. of H2, cyclohexane, 4 h) [312] and toluene (110 °C, 30 atm. of H2, cyclohexane, 4 h) [313]. The catalyst activity and selectivity were found to be strongly dependent on both the dendrimer generation and the ionic liquid nature, as well as on the reaction conditions used.
The main product of citral hydrogenation was citronellal (Scheme 8) [312]. The latter was found out to undergo cyclization to isopulegol—especially at longer reaction times, elevated temperatures, and decreases in the dendrimer generation, which authors explained by thermal decomposition of PAMAM dendrimers, taking place at temperatures of 100–150 °C [115,312,314]. The best results (the selectivity on citronellal of 100% at citral conversion of 50% after 4 h) were achieved for the catalyst, based on the dendrimer of the 5th generation, and under the use of [BMIM][NTf2] ionic liquid as catalyst coating [312].
As for partial toluene hydrogenation, the selectivity on methylcyclohexenes at 110 °C and 30 atm. of H2 in the presence of [EMIM][NTf2] ionic liquid did not exceed 12%, 21%, and 16% at conversion of 9–13% for the catalysts, based on PAMAM dendrimers of the 4th, 5th, and 6th generations, respectively (Figure 11a) [313]. Further, for the catalysts, based on PAMAM dendrimers of the 4th, 5th, and 6th generations, the selectivities on methylcyclohexenes at conversions of 69–73% under the same conditions reached 8.5%, 7.5%, and 6% respectively (Figure 11b) [313]. Thus, one can observe, that in the beginning of the reaction, the selectivity on methylcyclohexenes initially increases from the G4 to G5 catalyst and then decreases from the G5 to G6 catalyst (Figure 11). The authors explained the results obtained in terms of mean particle size, increasing with the increase in the generation of dendrimer [313,315].
It should also be noted that C=C double bonds more easily undergo hydrogenation in the presence of basic promoters [312,316]. The basicity of the PPI or PAMAM dendrimer increased with its generation and the number of terminal primary and node tertiary amino groups increase [317,318,319]. Hence, the results, obtained after partial hydrogenation of toluene in the presence of Ru nanoparticles, encapsulated into PAMAM dendrimers and supported on silica [313], appeared to be noticeably inferior to those obtained for conventional heterogeneous catalysts, especially containing the bimetallic RuZn phase, supported on the acid solid oxide, such as ZrO2 or CeO2 [315,320,321,322,323].

2.2. Impregnation/Adsorption on Carbon-Based Carriers

Yamamoto et al. developed catalysts based on DPA-encapsulated subnanoclusters, supported on graphitized mesoporous carbon (GMC), Ketjenblack (KB), and zirconia, for effective hydrogenation and oxidation reactions (Figure 12) [10,64,178,179,180,183,245,246]. Herein, carbon supports, such as GMC or KB, provided effective catalyst adsorption and stability due to the π–π interactions between aromatic fragments of dendrimer and carrier, thus preventing metal nanocluster post-aggregation [178,179,245].
In particular, Pt12 subnanoclusters with the mean diameter of 0.9 nm appeared highly active in the hydrogenation of various alkenes, including unreactive substrates, such as 4-chlorostyrene and, especially, 4-trifluoromethylstyrene [178], as well as superior poison tolerance to amines in the reductive hydroamination of aldehydes [179]—in contrast to 2.2 nm Pt nanoparticles, synthesized in the absence of the dendrimer template. TOF values up to 5850 h−1 in the hydrogenation of alkenes and quantitative yields of hydroamination products were achieved in the presence of the heterogeneous dendrimer-based catalyst Pt12@DPA-G4/GMC, already under the ambient conditions within 5 h (25 °C, 1 atm. of H2, MeOH) [178,179].
In the presence of tert-butylhydroperoxide under ambient conditions (25 °C, 6 h, n-decane) [245] or molecular oxygen (160 °C, 10 atm. of O2, 5 h) [180] DPA-encapsulated Pt subnanoclusters, immobilized on graphitized mesoporous carbon or Ketjenblack, afforded the oxidation of secondary alcohols, such as 1-phenylethanol [245] or aromatic hydrocarbons, such as toluene, respectively [180]. In the last case, the mixture of benzyl alcohol, benzaldehyde, and benzoic acid was formed, and the maximum TOF values reached up to 3300 h−1 [180]. It was established that oxidation, in both cases, proceeded via radical-chain mechanism, depending on cluster/particle size and metal nature [180]. In particular, larger sized nanoclusters (>30 atoms) promoted the cleavage of the O–O bond of the adsorbed O2 or tert-butylhydroperoxide molecule, thus providing the charge transfer to the adsorbed substrate molecule [180,245]. At the same time, electron-deficient subnanoclusters (12–19 atoms) of less oxophilic metals, such as Pt, tended to one-end coordination of O–O bond, thus providing the formation of peroxide radicals on the surface with the subsequent one-electron transfer followed by C–H bond cleavage, hence initiating toluene radical oxidation processes similar to homogeneous or nanoheterogeneous molecular oxidation catalysts [180,324,325]. More oxophilic metals, such as Ru or Cu, and vice versa, were prone to form partial surface oxides, resulting in a drastic decrease in the catalyst activity, with TOF values not exceeding 100 h−1 [180].
CunOx subnanoclusters, encapsulated in phenyl azamethyne dendrimers and immobilized onto ZrO2, appeared to be much more active catalysts under the same conditions [246]. The catalyst efficiency and TOF values, respectively, increased with the decrease in cluster size, reaching up to 720 h−1, which was due to the increased polarization of Cu–O bonds [246]. Nonetheless, the full conversion of toluene to benzoic acid was not completed, even after 128 h [246]. Introduction of additional metals, such as Pt and Au, in the cluster structure resulted in the essential increase in the catalyst activity in the oxidation of aromatic hydrocarbons (e.g., indane) by molecular oxygen, even under much milder conditions (90 °C, 1 atm. of O2, 6 h) [183]. The maximum TOF values, exceeding 600 h−1, were achieved for trimetallic CuPtAu clusters, that authors explained in terms of synergistic effects in alloy, arising from stabilization of Cu (0) and Cu (I) species by Pt, Ag, and Au, preventing the further oxidation of the former to Cu (II) [183,326,327]. It was established that oxidation of indane to 1-indanone proceeded mainly through the hydroperoxide stage, accompanied by Cu (0)/Cu (I) one-electron transitions on the interface between Cu, Au, and Pt [183].

2.3. Non-Covalent Immobilization on Modified/Coated Surfaces Using Hydrophobic Interactions or Hydrogen Bonding

Crooks and Stevenson suggested a method for the immobilization of PAMAM dendrimer-encapsulated monometallic Pd and Pt, or bimetallic PtPd nanoparticles on graphite or gold surfaces, using electrostatic interactions and hydrogen bonding [10,64,91,328,329]. For gold electrodes, α,ω-mercaptoundecanoic acid (MUA) was used as a non-covalent linker, and self-organization of near-surface charged monolayers took place (Figure 13) [26,91]. In the case of graphite electrodes, these interactions occurred between dendrimer terminal NH2 groups and -C(=O)OH groups, arising from defects on the graphite surface [328]. Hydrogen bonding provided the effective adsorption of PAMAM dendrimers, impregnated with Pt nanoparticles, on the surface of nitrogen-doped single-wall carbon nanotubes (SWCNT), synthesized by chemical vapor deposition (CVD) method (Scheme 9) [329]. Thus, obtained systems proved their efficacy in the electrocatalytic oxygen reduction, that could effectively proceed only for small nanoparticles (<2 nm in diameter), and the maximum activity was found for bimetallic nanoparticles Pt150Pd30, encapsulated into PAMAM dendrimers of the 6th generation, adsorbed on the graphite electrode [64,328].
Replacement of 11-mercaptoundecanoic acid with 11-mercaptoundecyl-1-tri(ethylene glycol) and 2-(11-mercaptoundecyl-1-hexa(ethylene glycol)) acetic acid, modified by 6-(maleic acid imide)hexanoic acid hydrazide, followed by covalent attachment to PAMAM dendrimers impregnated with Au nanoparticles, resulted in a high stable sensor for selective insulin detection with a concentration limit as low as 0.5 pM (Scheme 10) [62,330].
The more sophisticated approach was applied by Ghiaci’s group for the synthesis of bimetallic Mn/Co heterogeneous catalyst for p-xylene oxidation [324,331]. The negatively charged bentonite surface was treated here by cetyl pyridinium bromide; the hydrophobic tails of the latter retained the focal dodecyl moieties of PAMAM dendrons of the 1st generation, modified with salicylic aldehyde on the periphery (Figure 14) [331]. This bilayer system provided the location of Mn2+ and Co2+ ions, coordinated on the dendron salicylidene imine end groups, easily accessible for the p-xylene substrate. Hence, under the optimal conditions (KBr, AcOH, 190 °C, 17 atm. of O2, 3 h), a near to quantitative yield of terephthalic acid (97%) was provided. Nonetheless, in spite of the use of strongly coordinating chelating dendritic ligands, the catalyst synthesized appeared unable to effectively retain Mn2+ and Co2+ ions, resulting in the sharp downfall for the yield of terephthalic acid at recycling [331].

3. Covalent Immobilization on Heterogeneous Supports

Heterogeneous catalysts, based on PAMAM, PPI, poly(aryl ether), poly(ester amide), etc., dendrimers, and dendrons covalently attached to various preliminary functionalized insoluble supports, such as amorphous and mesoporous silica, polystyrene, carbon nanotubes, etc., have been developed since the end of 1990s [10,25,26,62,64,90,92,93]. Herein, dendrimer end group post-modification and metal deposition (both complex formation and nanoparticle encapsulation) is performed in the same way, as for homogeneous catalysts [26,146,147,148,149,171,193].
Covalent dendrimer/dendron grafting may be realized in both divergent (otherwise called “graft-from”) and convergent ways (otherwise called “graft-to”). The first one assumes the gradual, layer by layer, generation by generation dendron growth from the focal point, which is the functional group on the carrier surface (Scheme 11) [10,25,26,62,63,71,82,83,86,92,113,114,147,324]. The convergent approach, in its turn, implies the covalent attachment of the ready-synthesized dendrimers or dendrons to the preliminary functionalized support, and can be performed through the dendron focal point or dendrimer end groups [10,25,62,63,74,92,102,114,171,193]. Both of these approaches are further considered in the present review.

3.1. Divergent Dendrimer Grafting

3.1.1. Divergent Dendrimer Grafting on Silica Supports

Silica appears to be the conventional support for dendron grafting; as usual, it was preliminarily activated by (3-aminopropyl)triethoxysilane (APTES) treatment, followed by repeated divergent dendron growth (Scheme 12) [92,93]. Thus, Chung and Rhee developed heterogeneous catalysts, based on PAMAM dendrons of different generations, attached to silica through 3-aminopropyl spacer, and applied them in the enantioselective addition of diethylzinc to benzaldehyde (0 °C, toluene, 48 h) [92]. The catalyst chirality was provided by the decoration of dendron end groups with (1R, 2S)-ephedrine (Figure 15). It was found out that at low dendron density, the catalyst activity, selectivity, and enantiomeric excess (ee) increased from G0.5 to G3.5 dendrons, revealing the positive dendritic effect, and then dropped back, when passing to the G4.5 dendron. Herein, SiO2-G3.5 catalyst in its performance was just slight inferior to homogeneous one and could be recycles at least three times without significant loss in effectiveness [92].
Kawi et al. synthesized epoxidation catalyst, based on Mn (II) complexes with salicylidene-terminated PAMAM dendrons of the 0–4th generations, grafted to silica (Scheme 12) [252]. The yields of styrene oxide reached 20–75% after 1 h at 0 °C, using meta-chloroperbenxoic acid (m-CPBA) as an oxidant and N-methylmorpholine N-oxide as a co-oxidant in CH2Cl2 medium, rather than increasing with the increase in the dendron generation. Thus, the strong positive dendritic effect was observed. Nonetheless, the use of bulky substrates with an internal C=C double bond resulted in the noticeable decrease in the catalyst activity and reaction rate, which was typically for sterically crowded dendrimers of higher generations [60,91,117].
The similar approach for dendrimer immobilization was applied by Wu and Zhang to design hollow capsule material, impregnated with gold nanoparticles [26,333]. Herein, silica gel particles of ~260 nm in diameter, coated with grafted PAMAM dendrons of the 4th generation and already containing Au0 nanoparticles with the mean particle size of 2.3 nm, uniformly distributed over the material surface, were treated with poly(sodium 4-styrenesulfonate) and poly(allylamine hydrochloride), forming polyelectrolyte layers. Subsequent removal of silica core in the presence of HF resulted in hollow capsules of ~200 nm in diameter (Scheme 13). Thus, the synthesized catalyst proved its effectiveness in the 4-nitrophenol reduction, giving the quantitative yields already within 30 min at room temperature and was reused 5 times without significant loss in activity [333].
Alper, with co-workers, suggested the decoration of grafted PAMAM dendron periphery by palladium or rhodium phosphine complexes, as well as by palladium PCP pincer complexes, through the spacers of various lengths (C2, C6, C12) (Figure 13) [93,146,147,148,149,335]. Thus, synthesized catalysts were successfully applied in the selective hydrogenation of dienes and polyenes [148], oxidation of cyclic and terminal alkenes [147], hydroformylation [146,335], carbonylation [149,332,334,336,337], and cross-coupling reactions [338,339].
In particular, the total yields of alkenes in the hydrogenation of dienes in the presence of Pd phosphine complex catalysts, based on PAMAM dendrons, anchored to silica, reached up to 99% under the ambient conditions (25 °C, 1 atm. of H2, MeOH, 0.5–24 h) [148]. Herein, the total selectivity on alkenes as well as catalyst stability and recyclability increased with the increase in the dendrimer generation and spacer length (C2, C6, C12) between end amino and phosphine groups in the ligand (Figure 16, left) [148]. Hydrogenation mainly proceeded via 1,2-addition; nonetheless, the product distribution was strongly dependent on the substrate structure and, as a consequence, on the intermediate stability [148,288,289], and dienes, giving the higher portion of unsubstituted terminal alkenes, tended to undergo further hydrogenation to alkanes, whose portion may reach 1–33% at full conversion of initial diene (Scheme 14) [148].
Applied in the oxidation of terminal alkenes to methylketones, using tert-butyl hydroperoxide as an oxidant (55 °C, nonane), the catalyst SiO2-G0-Pd (i.e., containing only 3-aminopropyl linker without grafted dendron) gave yields of 27–42%, which decreased from 1-octene to 1-decene and 1-tetradecene, and 11% for cyclohexene within 24 h [147]. Herein, the strong dendritic effect was observed. An increase in the spacer length (Figure 16) from C2 to C12 allowed for a significant increase catalyst activity and made it possible to oxidize even bulky substrates, such as 5-vinyl-2-norbornene and 4-phenyl-1-butene [147]. It should be noted that the efficiency of these hybrid catalysts, based on prolonged dendrons, depended on the dendrimer generation, spacer length, and substrate size simultaneously, and positive dendritic effect was observed.
Rh complex catalysts, based on the same PAMAM dendron containing support, afforded the hydroformylation of various alkenes [146]. Conversions of 95–99% were achieved for various alkenes within 24 h at 65–75 °C, with a syngas pressure of 70 atm. (CO:H2 = 1:1) in CH2Cl2 medium. Herein, linear alkenes, such as 1-octene, tended to mostly give linear aldehyde, whereas styrenes, vinylnaphthalenes, and vinyl acetates (i.e., with substituents, giving −M effect to C=C double bond) gave branched products with selectivity often exceeding 90%. SiO2-PAMAM-Rh catalysts display a slightly negative dendritic effect (catalysts of G0—G1 generations effectively worked even at room temperature) and were reused several times without a noticeable decrease in activity [146].
Carbonylation of terminal alkenes in the presence of Pd phosphine complex catalysts, decorating the PAMAM dendrons of 0–4th generations, anchored to silica, in the alcohol medium (115 °C, 7 atm. of CO, MeOH, PPh3, p-TsOH, THF or toluene) resulted in the ester formation, with a predominance of linear product, even for styrene-type substrates [149]. Herein, this linear/branched ratio improved at catalyst recycling due to the leaching of the adsorbed less hindrance Pd complexes favoring the branched product formation [149].
Using otho-allyl- or vinylphenols and anilines as substrates under the similar conditions (100–140 °C, 7–43 atm. of CO, 0–36 atm. of H2, toluene, or CH2Cl2, 1,4-bis(diphenylphosphino)butane (dppb)), it was possible to obtain five-, six- and seven-membered ring lactones and lactams with yields up to 90% (Scheme 15a,b) [332,334]. Varying the reaction conditions (temperature, solvent, CO/H2 ratio, and pressure) for the materials, modified by Pd PCP pincer complexes (Figure 16, right), one could direct the reaction to the five- or seven-membered ring lactone formation with the selectivity of 82–91%, significantly exceeding that for a homogeneous non-dendritic Pd PCP complex catalyst [334]. Heterogeneous dendritic PCP pincer Pd catalysts had high stability at recycling, extremely increasing from G0 to G1 dendron generation, with the 7-membered lactone as a predominant product even under the conditions for 5-membered lactone formation in the 1st cycle [334].
In the presence of SiO2-PAMAM-Gn-(Pd(PR2)2)n+1 (n = 0–3, R = Ph, Cy) catalysts under carbonylation conditions (100 °C, 7 atm. of CO, MeOH, NEt3), various iodoarenes were effectively converted to corresponding arylcarboxylates with quantitative yields within 24–36 h [332]. In the case of carbonyls, assuming cycle formation (e.g., dibenzoxazepinone), there was also possible the use of bromoarenes under elevated temperatures (80/120 °C, 7 atm. of CO, NEt(i-Pr)2, toluene), with yields exceeding 90%, even for dimeric substrates (Scheme 15c) [337]. Catalysts, based on dendrons at least of the 1st generation, appeared highly stable under reuse; for substrates, such as 2-(2-iodophenoxy)aniline, the yields maintained to as high as 95% during the 8 cycles [338].
The same type of catalysts showed high activity in the Heck cross-coupling reactions for both iodo- and bromoarenes with styrene, butyl, or methyl acrylate with the predominant formation of trans-products (120–140 °C, Na2CO3, or NaOAc, DMF) [338,340]. The maximum yields of 85–99% were observed for the substrates with electron withdrawing groups (C(=O)OMe, NO2)), and were maintained when recycling the catalyst [341,342]. Herein, the SiO2-G0-PCP-Pd catalyst appeared more stable when reused [93].
Hagiwara’s group has developed a heterogenous hybrid catalyst, combining properties of both supported the ionic liquid catalyst (SILC) and dendritic catalyst [343]. Terminally alkylated PAMAM dendrons of the 3rd generation were anchored to nano-silica with a mean particle diameter of 10–12 nm, followed by the impregnation of Pd(OAc)2, solved in the ionic liquid. Under these conditions Pd complexes were reduced and formed dendron-encapsulated clusters (Figure 17) arranged into the raspberry-shaped superstructures [343].
Thus, the synthesized Pd-containing catalyst was successfully applied in the Suzuki cross-coupling reactions, using various bromo/iodoarenes and arylboronic acids as substrates under mild conditions (25–60 °C, K2CO3, H2O/EtOH). High biarene yields were favored by the use of phenylboronic acids with donor substituents and arylbromides or aryliodides with acceptor ones, and the quantitative yields for 4-bromoacetophenone, coupled with phenyl- or 1-naphthylboronic acid were achieved already within 30 min at room temperature [343]. It was established that the gradual decrease in activity of the hybrid dendritic SILC catalyst at recycling originated from Pd cluster aggregation, not from Pd leaching [343].
Li and Xu proposed a heterogeneous catalyst for Suzuki and Heck cross-coupling reactions based on poly(ethylene imine) (PEI) dendrons anchored to silica [341]. The latter was activated by (3-chloropropyl)trimethoxysilane grafting, followed by end-Cl atom replacement with diethanolamine according to the SN2 mechanism (Scheme 16). The material obtained, containing OH-terminated dendrons of the 1st generation, was further treated with thionyl chloride to replace terminal OH groups by Cl, and then the procedure was repeated until Cl-terminated dendrons of the 2.5th generation were propagated and then capped with 1,4-diazabicyclooctane (DABCO) (Scheme 16). Thus, the synthesized carrier, coated by grafted DABCO-terminated dendrons of the 3rd generation, was impregnated with PdCl2 solution in EtOH followed by a reduction in the presence of hydrazine in situ, resulting in Pd nanoparticles of 3–10 nm in diameter (Scheme 16) [341].
The synthesized SiO2-PEI-G3-DABCO-Pd catalyst revealed high activity Suzuki (for p-bromoanisole and phenylboronic acid) and Heck cross-coupling reactions (for p-iodo- or bromoanisole and styrene). By carefully adjusting the reaction conditions (80 °C, 0.5 h, K2CO3, EtOH/H2O for Suzuki cross-coupling, and 130 °C, 1–3 h, K3PO4, DMF/H2O for Heck cross-coupling), it was possible to achieve the quantitative yields for desired products (Scheme 17) [341]. It was established that under the said conditions, yields of 95–99% within 0.5 h in the Suzuki cross-coupling reactions in the presence of the SiO2-PEI-G3-DABCO-Pd catalyst were maintained when bromoarenes with para- and meta-, donor and acceptor substituents, as well as phenylboronic acid with donor para-substituents, were used as substrates. Replacing bromoarene for chloroarene resulted in the essential downfall in the catalyst activity; nonetheless, the yields of 77–89% were obtained within 24 h at 130 °C and increased catalyst loading for chloroarenes with acceptor para-substituents (NO2, CN) [341].
Heck coupling proceeded with quantitative yields within 1 h only for para-substituted iodoarenes and styrene. The replacement of iodoarene by bromoarene and, moreover, for chloroarene resulted in the noticeable decrease in the catalyst activity, making the reaction strongly sensitive to substrate structure and electronic properties [344]. The yields of 60–89% within 2–5 h appeared to be achievable for bromo- and chloroarenes only with acceptor para-substituents (NO2, C(=O)OCH3), while the use of donor- or meta-substituted bromoarenes (bromobenzene, 4-bromotoluene, and anisole, 3-bromonitrobenzene) led to the sharp downfall in conversion to 0–25%. The use of acrylic acid, such as alkene instead of styrene, resulted in the yields of 83–84%, achievable only after 10 h both for iodo- and bromoarenes, which may be connected with a pH decrease in the reaction medium [341]. The hybrid SiO2-PEI-G3-DABCO-Pd catalyst appeared to be easily recoverable from the reaction mixture. Nonetheless, its activity noticeably decreased in the recycling of Suzuki coupling, which may be attributed to the gradual Pd leaching and/or Pd nanoparticle agglomeration [341].
Mohammadpoor-Baltork and Mirkhani developed Pd nanoparticles, immobilized on nano-silica (10–40 nm), coated with triazine dendritic polymer (nSTDP) [342,345]. The latter was propagated by the repeated alternating stages of triazine trichloride (cyanuric chloride) and bis(3-aminopropyl)amine addition according to the Csp2 SN2 mechanism, until G3 dendrons were obtained (Scheme 18). The mean particle size of the further deposited Pd nanoparticles was found to be 3.1 nm [342,345].
Thus, the synthesized hybrid Pd-nSTDP catalyst was studied in Heck and Suzuki C–C cross-coupling reactions (Scheme 19) [342], as well as in C–S cross-coupling reactions between various arylhalides and diaryldisulfides under both microwave (MW) and conventional thermal conditions (Scheme 20) [345]. Suzuki coupling proceeded with near quantitative yields (93–96%) within 2–3 h or 6–10 h at room temperature for iodoarenes or bromoarenes, respectively (K2CO3 as a base, DMF/H2O (1:3) as a reaction medium) (Scheme 19) [342]; the use of MW irradiation (200 W, 70 °C) allowed the achievement of the same yields after 2–5 min, both for iodoarenes and bromoarenes. Chloroarenes required much longer reaction times (18–24 h; 7–10 min under MW) or more severe reaction conditions (80 °C). Heck coupling between various arylhalides and substituted styrenes required 9–20 h for the near-to-quantitative yields (90–95%) to be reached under the conventional thermal conditions (85 °C, K2CO3, DMF/H2O (1:3)), and 8–20 min under MW conditions (200 W, 70 °C), depending on the halide nature [342]. The presence of electron-withdrawing or, especially, of electron-donating groups in the structure of both substrates resulted in the noticeable decrease in the reaction rate, as in Suzuki and as in Heck coupling reactions [342].
In C–S coupling (Scheme 20), yields of 90–95% were achieved within 5 h under conventional thermal conditions (80 °C, Bu4NOH, DMF/H2O (1:1)) or within 10–15 min under microwave (MW) conditions (Scheme 20) with the TOF values, reaching, in the last case, up to 5700 h−1, enormously exceeding those for conventional homogeneous and heterogeneous catalytic systems, such as PdCl2(dppf)/Zn, CuFe2O4 or Cu2S/Fe [345]. It was found that electron withdrawing para-substituents favored the better reaction yields; nonetheless, it was noticeable only for chloroarenes, converted to the corresponding diarylsulfides with yields of 90–90% after 10 h. Moreover, double C–S coupling was successfully conducted with the various para-and meta-dibromoarenes (including 2,6-dibromopyridine) and diaryldisulfides (including 2,2′-dithiobis(benzothiazole)) under said conditions. The yields appeared as high as 75–95% within 6–10 h under conventional thermal heating and within 10–45 min under MW irradiation. As in the case of mono C–S coupling, the reaction rate and arene disulfide yields were mainly enhanced by electron withdrawing groups (EWG: phenyl, benzo, C=N/pyridine) in the substrates [345]. Hence, using the high activity of a Pd-nSTDP catalyst even for chloroarenes, especially under the MW conditions, it was possible to synthesize via C–C or C–S coupling reactions with banana- and star-shaped molecules with benzene, pyridine, or triazine cores [342,345].
The hybrid dendritic Pd-nSTDP catalyst appeared highly stable at recycling, and no decrease in the reaction rate and product yield, nanoparticle agglomeration, or noticeable Pd leaching were observed after seven runs in both C–C and C–S coupling reactions under both thermal and MW conditions [342,345]. A reaction mechanism was proposed, and accordingly, dendron moieties functioned as micro/nanoreactors for Pd nanoparticles, strongly retaining them and, therefore, prevented their leaching (Scheme 21) [344,345,346,347,348]. Herein, at the first stage, an oxidative addition of arylhalide (ArX) to the surface of Pd nanoparticle took place, followed by diaryldisulfide addition and halide release in the presence of base. Re-coordination of chemisorbed aryl and aromatic sulfide resulted in the reductive elimination of the end product, diarylsulfide, and the subsequent turn-back of Pd species to the next cycle (Scheme 21) [345].
It is important to note that dendron grafting on the silica surface is sufficiently intricate, with the increase in the dendron generation and steric hindrances arising from this and resulting in the incomplete Michael addition and amidation [92,146,252,339,349,350]. Herein, the yields of G3–G4 and higher dendron grafting may not exceed 40–60% [147,252,339,350]. In order to improve the performance of silica-supported dendritic catalysts, it was suggested by Alper and Cao to attach future dendron focal points to the inner walls of the ordered mesoporous materials, such as pore-expanded MCM-41 [351] and SBA-15 (Scheme 22) [352,353].
In particular, it was established that the yield of G3–G4 PAMAM dendrons (up to 4.5 m in diameter [85,114,354,355,356]), anchored to pore-expanded MCM-41 material with channel diameter of 10.6 nm, reached 70–75% [351], which was also due to the lower initial APTES loading in comparison with the grafting to amorphous silica [147,339,350], hence resulting in the lower dendron density and less steric hindrances. Analogously covalent immobilization of G3–G4 PAMAM dendrons in the pores of SBA-15 material (pore diameter of ~7.5 nm) proceeded with the yield of ~60% [350].
The last material was further used as a carrier for monometallic Pt, Pd, Au, and bimetallic PdAu and PdPt nanoparticles, resulting in the catalysts for hydrogenation of C=C double bonds [353], reduction of p-nitrophenol [354], oxidation of alcohols [355], and Suzuki cross-coupling [357]. Herein, the presence of covalently grafted dendrons provided uniform and narrow well-dispersed particle size distribution (1–3 nm) inside the SBA-15 channels [352,353,357]. Thus, Pd-containing catalysts with the particle size of 1.2–3.4 nm, based on the dendritic organic-inorganic hybrid composites, revealed extremely high activity in the hydrogenation of allyl alcohol under the ambient conditions (30 °C, 1 atm. of H2, MeOH), reaching conversions of more than 99% and TOF values of 710–2270 h−1 within 20–120 min [353]. A noticeable negative dendritic effect was observed, appearing as a decrease in the reaction rate for higher generation catalysts; nonetheless, the catalyst based on dendrons of the 3r generation mostly suppressed isomerization of allyl alcohol to acetone and demonstrated the best selectivity on 1-propanol, reaching up to 95% [353]. Moreover, the catalyst SBA-15-PAMAM-G3-Pd revealed the highest stability at recycling and storage and was successfully reused several times without any loss of activity, Pd leaching, or nanoparticle agglomeration observed [353].
A similar approach can be applied for the synthesis of G4-PAMAM dendron-encapsulated monometallic Pd and bimetallic PdAu nanoparticles, immobilized in the SBA-15 channels by co-complexation method [357]. These dendrimer-encapsulated PdAu nanoparticles effectively catalyzed Suzuki cross-coupling between various arylbromides and phenylboronic acid under the microwave conditions (100 °C, 30 min, K3PO4, H2O/EtOH, MW), with the yields significantly exceeding those for monometallic Pd catalysts, active mostly in the iodoarene coupling. This phenomenon was attributed to Au3+ ions weakening the coordination interactions between Pd2+ and dendron amino groups, hence facilitating the reduction of Pd2+ to Pd0 [357]. The maximum yields (90–99%) for both Pd and PdAu catalysts were achieved using unsubstituted substrates or para-substituted phenylboronic acids with electron donating groups (CH3, OCH3) and para-substituted arylhalides with electron withdrawing groups (NO2). Herein, the use of sterically hindered ortho-substituted phenylboronic acids, even coupled with iodebenzene, resulted in a drastic decrease in the reaction yields down to 30% due to the negative dendritic effect [358]. It was discovered that the mean particle size, arising from the initial metal/dendrimer ratio, had an essential effect on the reaction rate and cross-coupling yield, and the best activity and stability at recycling were revealed for tiny Pd10 and Pd5Au5 nanoparticles of 1.4–1.6 nm in diameter strongly encapsulated inside the dendron cavities [357].
Monometallic Au nanoparticles, encapsulated in PAMAM dendrons of G1–G4 generations, grafted to SBA-15 ordered mesoposous silica, effectively catalyzed the oxidation of secondary alcohols at room temperature and the ambient oxygen pressure in biphasic CH2Cl2/H2O reaction medium, in the presence of K3PO4 as a base [355]. An increase in the dendron generation from G1 to G4 was found to increase mean particle size from 1.6 to 3.4 nm, which authors explained in terms of steric crowding on the higher generation dendron periphery, interfering the Au3+ ions to penetrate inside the dendron cavities, thus resulting in the outer dendron-stabilized nanoparticles [355]. The reaction rate and the yield of corresponding ketone decreased with an increase in the dendrimer generation and increased with a decrease in mean particle size arising from the Au/dendrimer ratio. In particular, the yield of 99% for acetophenone in the presence of the SBA-15-PAMAM-G4-Au100 catalyst was attained only within 24 h. Rather, it decreased down to 80–90% for bulky or negatively induced substrates, such as 1-(2-naphthyl)ethanol or 1-(4-chlorophenyl)ethanol, which are able to competitively adsorb via their substituents, or, especially, replace aromatic by alicyclic substrates, such as cyclopentanol or cyclohexanol (50–70%); for primary linear aliphatic C6–C8 alcohols the conversions did not exceed 4–20% [353]. Nonetheless, in the oxidation of aromatic alcohols, the SBA-15-PAMAM-G4-Au100 catalyst can be recycled several times without noticeable loss in activity—in contrast to Au nanoparticles, which are supported on SBA-15 without a grafted dendrimer ligand [353].
With respect to heterogenized G4-PAMAM, dendron-encapsulated Pt nanoparticles, it was discovered that increase in mean particle size (at elevated Pt loadings) resulted in the decrease in the reaction rate both for p-nitrophenol and K3[Fe(CN)6] reduction (NaBH4 or Na2S2O3 as a reducing agent, 25 or 40 °C, H2O, 30 or 90 min, respectively) [354]. It was presumably due to larger Pt nanoparticles blocking the channels of SBA-15, hence reducing the catalyst-specific surface area and retarding the mass transfer in the reaction medium [354]. Nonetheless, in contrast to SBA-15-PAMAM-Pd catalysts [354], the analogous Pt-containing catalyst appeared did not seem very stable at reuse, maintaining about 50% of initial activity to the fifth cycle due to mechanical losses [354].
Shantz et al. developed monometallic Cu and bimetallic CuAu catalysts, based on Simanek-type melamine-piperidine dendrons, covalently grafted inside SBA-15 channels (Scheme 23) [233]. Herein, in spite of the fact, that bimetallic catalyst was synthesized by co-complexation method, Au core–Cu shell nanoparticles were obtained. The catalysts obtained were studied in the alkyne-azide click cycloaddition (50 °C, NEt3, THF), and strong negative dendritic effect was observed—also, in comparison with a conventional heterogeneous catalyst Cu/SBA-15 without a dendrimer. The yields of 94% and 81% were reached for Cu/SBA-15 and Cu@G2/SBA-15 after 5 and 24 h, respectively; nonetheless, the presence of dendrimer was found to strongly suppress metal leaching at recycling. The presence of gold in the nanoparticle structure provided a remarkable increase in the catalyst activity: a 100% conversion Cu@G2/SBA-15 catalyst was achieved after 15 h with the reaction rate, quite similar for that of the Cu/SBA-15 catalyst.
By means of IR spectroscopy, it was established that alkyne adsorbed rapidly on the copper surface without any induction period before azide adsorption, thus assuming the linear intermediate formation, antecedently the ring closure and triazole formation, to be the rate-limiting step in the reaction (Scheme 24) [233]. Herein, all intermediates observed were the same, as in homogeneously catalyzed alkyne-azide cycloaddition. No electron synergistic effect between copper and gold was revealed, indicating that the increase in activity for the bimetallic catalyst was mainly due to enhanced copper atom distribution on the nanoparticle surface [233].

3.1.2. Divergent Dendrimer Grafting on Polymer and Carbon Supports

Linear insoluble polymers, such as polystyrene (PS), are another common support for dendron grafting, thus resulting in dendronized polymers (DenPols), which can be synthesized in different ways depending on the dendrimer’s nature [63,72,75,86,93,358,359,360,361]. In particular, to produce a focal point of future PAMAM dendrons supported on polystyrene, the latter is preliminarly activated by a chlorometylation reaction followed by substitution for the amino group directly according to the SN2 mechanism [362,363] or via phthalimide stage (Scheme 25) [364,365,366,367].
Thus, Feng et al. studied the hydrogenation of 5-membered heterocycles, such as furan, pyrrole, and thiophene, in the presence of Pd catalysts, based on salicyliden-terminated PAMAM dendrons of the 2nd generation (similar to those in Scheme 12), anchored to polystyrene (PS), at an ambient temperature and pressure in the ethanol medium [368]. Quantitative conversions were achieved for all substrates, and products of ring opening and C–C bond cleavage were also obtained. The PS-PAMAM-Pd catalyst was successfully recycled 10 times with a gradual decrease in activity [368].
Krishnan and Sreekumar used polystyrene-supported PAMAM dendrons of various generations for the synthesis of encapsulated Pd nanoparticles, followed by their application in the Suzuki cross-coupling (100 °C, Na2CO3, 1,4-dioxane/water) [364]. Herein, Pd nanoparticles, stabilized by G1 and G2 dendritic polymers, tended towards agglomeration, whereas G3 support provided well-separated nanoparticles of 10–20 nm in diameter (Figure 18). In the presence of a catalyst, based on dendrons of the 3rd generation, the reaction between phenylboronic acid and 4-nitroiodobenzene proceeded with the yield of 90–95% within 12–24 h. It was established that the use of substrates with donor substituents (CH3, C(CH3)3, NH2) both for arylboronic acid and, especially, for arylhalide resulted in the significant decrease in the catalyst activity and reaction yield, which was extremely noticeable for arylbromides and arylchlorides. PS-PAMAM-G3-Pd had high stability at recycling, and only a slight decrease in the reaction yields was observed during 6 cycles due to Pd leaching [364].
The analogously hybrid metal complex catalyst Mn (II), based on PS-PAMAM-G3 dendritic carrier, demonstrated a high efficiency in the oxidation of various secondary alcohols, giving near-to-quantitative conversions (94–98%) within 3 h at room temperature, using K2Cr2O7 or KMnO4 as oxidizers and acetone as solvent [252]. PS-PAMAM-G3 material itself effectively catalyzed Knoevenagel condensation, giving yields of 95–100% within 10–120 min at 30–50 °C, using ethanol as solvent [366]. Herein, dendritic material acted as a base catalyst, and the reaction rate decreased with the increase of both carbonyl and the active methylene compound [366]. For both above-mentioned materials, a strong, positive, dendritic affect and excellent recyclability were revealed [252,366]. An increase in the degree of the polystyrene cross-linking furthered a slight conversion decrease [252], whereas an increase in the dendronization degree, in comparison, positively influenced the catalyst activity [366].
Jiang, Yang, and Pan suggested a novel approach for the synthesis of gold nanoparticles, stabilized by PAMAM dendrons divergently grafted to polystyrene beads [367]. This approach included preliminary protonation of alkylated tertiary dendron end groups, followed by complex formation with NaBH4; this complex functioned as a reducing agent for KAuCl4, added after all the others, thus resulting in nanoparticles of 5–10 nm in diameter (Scheme 26).
Thus, synthesized polymeric catalysts were studied in the oxidation of primary and secondary benzyl alcohols and in arylboronic acid homocoupling (Scheme 27). As for primary benzyl alcohols, the yield of corresponding benzaldehydes reached 99% within 16 h for substrates with +I or +M para-substituents (Me, OMe, Cl, Br, iPr) at room temperature and ambient oxygen pressure in the presence of the PS-PAMAM-G3-Au catalyst [367]. For the secondary alcohols, the quantitative conversions to corresponding methylketones were achieved only within 32 h, all other conditions being equal. Replacement of the substituent in para-position with the hydroxymethyl group by the NO2 group’s strong −M-effect resulted in a drastic decrease in the reaction yield down to 13%.
Additionally, it should be noted that despite the PS-PAMAM-G3-Au catalyst exhibiting the best efficiency, for all other catalysts, based on lower generation dendrons, a strong negative dendritic effect was observed, and the yields of benzaldehyde reached 73%, 50%, and 39% for G0, G1, and G2 catalysts, respectively [367]. Homocoupling of phenylboronic acids in the presence of the PS-PAMAM-G3-Au catalyst proceeded with the yield of 94–99% within 24 h at room temperature, air subjected, in water medium, using K2CO3 as a base. Herein, biphenyls were selectively produced from the substrates with donor meta- and para-substituents (Me, OMe), whereas ortho-substituted substrates resulted in hydrolysis product (phenol) [367]. The catalyst PS-PAMAM-G3-Au appeared very stable at reuse and maintained a yield in benzyl alcohol oxidation as high as 99% after 16 h and 52–55% after 5 h during 14 cycles [367].
Arya, Alper, and Manzer designed polystyrene-supported poly(aryl amide) dendrons in a divergent route, using a peptide synthesis approach (Scheme 28), and then modified them by Rh phosphine complexes [369]. The synthesized heterogeneous Rh complex catalysts were applied in the hydroformylation of various olefins at 65 °C and 140 atm. of syngas (CO:H2 = 1:1). The positive dendritic effect was discovered, and quantitative conversions were achieved within 24 h for G2–G3 catalysts, which did not lose their efficacy for 4–6 cycles at reuse. Herein, the portion of branched product often exceeded 90% not only for styrene-type substrates, but also for vinyl acetate [369].
Dahan and Portnoy suggested materials, based on Fréchet type poly(aryl ether) dendrons, immobilized on hydroxylated polystyrene (Wang resin) [370,371]. The dendron grafting was proposed here in two different ways (Scheme 29) [370,371]. The first one suggested the attachment of 5-hydroxyisophthalate moiety to hydroxylated polystyrene via a Mitsunobu reaction in the presence of diisopropyl azodicarboxylate (DIAD) and triphenylphosphine-sulfonamide betaine in CH2Cl2 medium, followed by reduction using a LiBH4 reaction in the presence of B(OMe)3 in THF medium (Scheme 29a) [372]. According to the second approach, 3,5-bis(acetoxymethyl)-phenol unit was grafted to Wand resin via a Mitsunobu reaction again in the presence of PPh3 and DIAD in THF medium, followed by hydrolysis of ester groups in the presence of tetrabutylammonium hydroxide (TBAH) in water/THF medium (Scheme 29b) [373]. Thus, hydroxymethyl-terminated poly(aryl ether) dendrons of various generations were obtained; their subsequent modification by diphenylphosphinobenzoic acid in the presence of diisopropylcarbodiimide (DIC) in CH2Cl2 (for ortho-isomer) or DMF medium (for para-isomer), catalyzed by 4-(N,N-dimethylamino)pyridine (DMAP), resulted in the heterogenized diphenylphosphine-terminated ligands for metal complex catalysis [370,374].
In particular, Co-containing catalysts proved their efficacy in Pauson–Khand [2+2+1] cycloaddition of alkyne, alkene, and carbon monoxide (Scheme 30) [374], whereas Pd-containing catalysts appeared highly active in the Heck cross-coupling reaction (Scheme 30) [370]. A positive dendritic effect was found out in both processes, which was explained in terms of a decrease in the resin internal coordination cross-linking, an increase in the number of catalytic sites on the polymer surface, and enhanced stability of the metal nanoparticles formed in situ when the dendron generation increased [370,374]. The best yields in Pauson–Khand cycloaddition were as high as 68% within 24 h at 70 °C and an ambient CO pressure in THF medium, using Co-containing catalysts with o-diphenylphosphinobenzoyl dendron end group moieties, which authors attributed to the Co bidentante coordination [374].
In the Heck coupling between bromobenzene and methyl acrylate (Scheme 31), the quantitative yield of methyl trans-cinnamate was analogously obtained for the catalyst, based on dendrons of the 3rd generation, decorated on periphery by Pd complexes with p-diphenylphosphinobenzoyl fragments, rather than exceeding that for the homogeneous catalyst under the same conditions (80 °C, NEt3, NMP, 72 h) [370]. Herein, milder reaction conditions (80 °C, 72 h vs. 120 °C, 14 h) provided much higher selectivity on cross-coupling products in comparison with bromobenzene homocoupling (200:1 vs. 32:1). Replacement of methyl acrylate by styrene, possessing less −M-effect, required more severe reaction conditions for better conversion and selectivity, whereas +M butyl vinyl ether resulted in a mixture of products of both α- (α-butoxystyrene and acetophenone) and β-arylation (cis- and trans-β-butoxystyrenes), and gradually increased together with the increase in the dendron generation along with the β/α ratio, which decreased. The phenomenon observed was explained in terms of the higher portion of the olefin insertion proceeding through cationic intermediates, arising from +M-effect of the OBu group and increasing in the polar environment, provided by the catalytic species of a higher generation [370].
Other interesting carbon materials for dendron immobilization are carbon nanotubes (CNTs) [372] and polyaromatic frameworks (PAFs) [363]. The latter (otherwise called porous aromatic frameworks) are three-dimensional polymers, who’s rigid, diamond-like structure, consisting of tetrahedral node atoms connected with polyphenylene edges (Figure 19), provides high specific surface area and thermal stability [375,376,377,378]. Earlier it was demonstrated that Pd nanoparticles, immobilized on polyaromatic frameworks, functionalized with amino groups, appeared noticeably inferior to those based on unmodified PAFs, as catalysts for hydrogenation of various unsaturated compounds (phenylacetylene and styrene, linear alkenes, alkynes, and dienes) [379]. Nonetheless, this Pd/PAF-NH2 catalyst exhibited excellent stability, maintaining a conversion of about 30% in the hydrogenation of phenylacetylene during 6 cycles [379]. Herein, amino groups attached to polyphenylene edges of PAFs can be considered as G0 dendrons.
The dendron grafting to PAF was performed via chloromethylation stage followed by replacement with diethanolamine according to SN2 mechanism (Scheme 25) [363]. Thus, polyaromatic frameworks were obtained, coated in OH-terminated PEI dendrons of the 0th generation. Subsequent chlorination with thionyl chloride and replacement with diethanolamine again gave materials with dendrons of the 1st generation (Scheme 32) [363]. The deposition of Pd nanoparticles was conducted in a typical way for polyamine dendritic catalysts, including the impregnation with Pd(OAc)2 solution in chloroform followed by the reduction in the presence of sodium borohydride in the ethanol/water mixture [363]. It was surprising that more hindered G1 dendrons favored larger particle formation with broader-sized distribution in comparison with less hindered G0 dendrons (4.5–13 nm vs. 1.5–4 nm), which authors attributed to the blocking of pores by branched aminoethanol moieties and, as a consequence, nanoparticle formation outside the pores [363].
The grafting of diethanolamine moieties to a PAF surface significantly enhanced the catalyst performance in the hydrogenation of various unsaturated compounds (cyclic and linear alkenes, alkynes and dienes, phenylacetylene, and styrenes) in comparison with just amino-modified PAF, allowing it to reach TOF values more than 100,000 h−1 versus ~55,000 h−1 for material based on PAF-20 with the terphenylene edges [363,379]. It should be noted that the activity of PAF-based catalysts is strongly dependent on the pore size determined by the number of phenylene rings in the edges [377,378]. As a consequence, change from Pd-PAF-20-PEI-G0 to Pd-PAF-30-PEI-G0 with quarterphenylene edges resulted in the increase in the catalyst-specific activity up to 300,000 h−1 for linear alkynes and dienes (Figure 19) [363]. Dendron grafting and further propagation decreased the pore volume and specific surface area; as a consequence, the specific catalyst activity decreased again, and conversions often did not exceed 1–5%. Nonetheless, for 1-hexyne the TOF values of as high as 100,000 h−1 and 190,000 h−1 for Pd-PAF-20-PEI-G1 and Pd-PAF-30-PEI-G1, respectively, were achieved (Figure 20) [363]. Similar to Pd-PAF-NH2 [380], Pd-PAF-PEI catalysts proved their stability at recycling. An observed decrease in activity at the first two cycles, especially noticeable for Pd-PAF-20-PEI-G0, was attributed to the leaching of Pd nanoparticles formed outside the pores [363].
With regards to carbon nanotubes, their functionalization is mainly performed via oxidazing treatment by an HNO3/H2SO4 mixture, causing structure defects on the surface (C(O)OH, C=O and OH groups, arising from C=C and C–C bond cleavage) [372]. In particular, by the esterification of the surface carboxyl groups, arising from the defects on the outer carbon nanotube walls, with propargyl alcohol followed by the cycloaddition of 3-azidopropylamine in the presence of Cu(OAc)2 and sodium ascorbate as catalysts, the focal points (G0) for the subsequent PAMAM dendron grafting were produced (Scheme 33) [372]. The latter was conducted by the divergent method, using the repeated stages of methyl acrylate Michael addition and the subsequent amidation by ethylene diamine (Scheme 33). Hence the materials, based on multi-wall carbon nanotubes coated with dendrons of up to the 3rd generation (PAMAM-g-MWCNTs), were obtained and used for Pd nanoparticles deposition analogously to the homogeneous PAMAM dendrimer-based catalysts [115,372]. Mean particle size was found to be 1.8–2.8 nm for Pd-G3-PAMAM-g-MWCNTs material.
The synthesized Pd-PAMAM-g-MWCNTs catalysts were examined in the Heck cross-coupling reaction, and a strong positive dendritic effect was revealed, which authors explained in terms of higher Pd nanoparticle stability and leaching retarding in the presence of dendrimers of higher generations [372]. Yields of 95–99% were achieved already after 1–6 h at 100 °C, with the presence of K2CO3 as a base in the NMP medium, even for chloroarenes. Herein, electron withdrawing substituents (such as NO2 or C(=O)OMe) in both arylhalide and alkene favored the improved catalytic activity and, therefore, to the maximum reaction yields. Pd-G3-PAMAM-g-MWCNTs revealed the best stability at recycling, and no decrease in the catalyst performance was observed during six cycles [372].

3.2. Convergent Dendrimer Grafting

The convergent approach, implying the covalent attachment of the already synthesized dendrimers or dendrons to the preliminary functionalized carrier, escapes the defects resulting from irregular dendrimer coatings, which are inevitable for multi-step divergent synthesis due to diffusion limitations increasing with the increase in the dendron generation [92,146,147,252,339,349,350,364,368,381]. The anchoring of readily prepared dendrimers of the appropriate generation can be performed in one stage, thus providing an atom-economical strategy and, as a consequence, much higher yields of the dendronized hybrid support [382].

3.2.1. Convergent Dendrimer Grafting to Carbon-Based Supports

Giacalone, Gruttadauria, and Prato have developed materials, based on PAMAM dendrons of the 2.5th and 3rd generations, covalently attached to single wall carbon nanotubes (SWCNT) in one step (Scheme 34) [10,64,382]. Initial PAMAM dendrimers with a disulfide core underwent radical cleavage on S–S bonds, easily grafting after to non-functionalized pristine nanotubes [380]. The subsequent treatment with Na2PdCl4 aqueous solution followed by NaBH4 reduction resulted in well dispersed palladium nanoparticles with mean sizes of 3.4 and 1.6 nm for G2.5- and G3-based catalysts, respectively.
The catalysts obtained, based on dendrons of the 3rd generation, revealed extremely high activity in Suzuki and Heck cross-coupling reactions, providing quantitative yields within 5–22 h under moderate conditions (Suzuki coupling: ArBr, Ar’B(OH)2, K2CO3, EtOH/H2O (1:1), 50 °C, 5–22 h; Heck coupling: ArI, H2C=CH–C(=O)OMe, NEt3, DMF/H2O (4:1), 90 °C, 5–16 h) (Scheme 35) [382]. Herein, in Suzuki coupling, the maximum reaction rate and highest product yields were observed for unsubstituted phenylboronic acid and arylbromides with EWG substituents in para-position; moreover, a SWCNT-PAMAM-G3-Pd catalyst appeared to be able to promote coupling for aryldibromides and even for arylchlorides. Heck coupling of methyl acrylate in the presence of a SWCNT-PAMAM-G3-Pd catalyst proceeded similarly for aryliodides with different substituents (in para and meta positions, +I, +M, –M), giving yields of 96–99% within 5 h—except 4-nitrophenyliodide, for which an 85% yield was achieved only after 16 h.
The catalyst SWCNT-PAMAM-G2.5-Pd, based on the ester-terminated dendrons of the 2nd generation, appeared much less active and much more sensitive to substrate structure and electron properties, providing the yields of 97–97% only after 22 h, which authors attributed to the worst stabilization of Pd nanoparticles by ester groups in comparison with amino groups [382]. Under microwave conditions at 120 °C in the presence of the SWCNT-PAMAM-G3-Pd catalyst, it was possible to reach the quantitative yields already within 5 min, thus providing TOF values up to 550,000 h−1. This extremely high catalyst activity was explained in terms of the “release and catch” mechanism (Scheme 36) [382,383,384] according to that Pd species, interacting with substrate molecules, leached from the nanoparticle surface, and, therefore, the catalytic cycle partly proceeded in the homogeneous phase; on the other hand, the released Pd species can be captured by the dendron moieties, forming nanoparticles again, but of a larger size (3.5 nm). Nonetheless, in spite of partial Pd leaching in the process, the catalyst SWCNT-PAMAM-G3-Pd could be easily recovered and maintain its efficiency in cross-coupling reactions, thus providing yields as high as 99% during 6 cycles at least [382].
Earlier, the anchoring of poly(aryl ether) Fréchet type dendrons to preliminary oxidized multiwall carbon nanotubes (Scheme 37) [86,385] and poly(ester amide) Newkome type dendrons (Scheme 38) to single-wall carbon nanotubes [25,62,386] were described. In the last case, the material obtained was used for immobilization of CdS nanoparticles as small as 1.4 nm in diameter to develop very stable quantum dots (QD), maintaining their UV absorption behavior more than 3 months—in contrast to CdS nanoparticles, stabilized by ungrafted dendrons and, therefore, tend towards aggregation and precipitation [25,386,387,388].
Bimetallic PtRu nanoparticles of 2.6 nm in diameter, prepared by co-complexation method and encapsulated into PAMAM dendrimers covalently grafted to oxidized carbon nanofiber (CNF, Scheme 39), proved their efficacy in the electrocatalytic methanol oxidation, demonstrating a sufficiently better performance in comparison with commercial PtRu/C catalyst [389].
Another strategy was developed for the anchoring of PPI dendrimers to the inner walls of the ordered mesoporous resins (OMR) [362,390]. The latter are carbon materials of SBA-15 or MCM-41 type, obtained by a soft-template method from the polymerization of phenol and formaldehyde in the presence of a poly(ethylene glycol)-block-poly(propylene glycol)-block-poly(ethylene glycol) template (Pluronic P123 or F127) [391]. Anchoring was conducted through the chloromethylation stage for the activation of the carrier surface (Scheme 40) [362,390]. Further deposition of Pd nanoparticles was performed in two-stage method, typical for PAMAM and PPI dendrimers and including complexation with Pd2+ ions, followed by reduction in the presence of NaBH4 [362,390]. The mean size was 3.7 nm, which was consistent with the pore size of MPF-PPI-G3 material (Figure 21) [362,390].
Thus, obtained hybrid OMR-PPI-G3-Pd catalyst revealed a good performance in the hydrogenation of naphthalene (400 °C, 50 atm. of H2, 5 h, hexane) [362] and phenylacetylene (80 °C, 10 atm. of H2, 15–60 min, solventless) [362], and an excellent one in the hydrogenation of 1-octyne, 4-octyne and 1-hexyne (80 °C, 10 atm. of H2, 15–120 min, solventless) [390]. Hydrogenation of octynes proceed with a quantitative conversion already within 15 min, giving TOF values up to 2000 min−1 (~120,000 h−1) and total selectivity on alkenes of ~98%, with the isomerization portion for octenes not exceeding 10–15% [390]. Replacement of 1-octyne for tinier 1-hexyne resulted in a noticeable decrease in both catalyst activity and selectivity (conversion of 99% within 60 min, total selectivity on alkenes of <90%, TOF ~ 1500 min−1/90,840 h−1), which authors attributed to the worse adsorption of 1-hexyne [392]. Replacement of linear alkynes by bulky phenylacetylene required 2 h for quantitative conversion (TOF ~1000 min−1/61,920 h−1), while selectivity on styrene was 92% [362].
Hydrogenation of naphthalene, in the presence of the same catalyst, resulted in the preferential decalin formation with the selectivity of 81% at a conversion of 90% within 5 h (TOF ~ 220 min−1/13320 h−1) [362]. The major product of guaiacol hydrogenation was 2-methoxycyclohexanol with the selectivity of 90–97% (100 °C, 50 atm. of H2, 1–5 h, H2O) [393]. Addition of sulfuric acid to the said catalytic system, as well as prolongation of the reaction time, favored the increased portion hydrodeoxygenation products, such as methylcyclohexane, methoxycyclohexane, and cyclohexanol, with the total selectivity being up to 98% at 100% guaiacol conversion, thus significantly exceeding those for a conventional Pd/C catalyst [394]. The OMR-PPI-G3-Pd catalyst did not lose its activity and selectivity at recycling in the 1-octyne hydrogenation [390], maintaining a conversion of more than 95% during 6 runs; however, it appeared unstable under the conditions of guaiacol [394] and, especially, of naphthalene hydrogenation [362].
Another interesting example of a “graft-to” convergent strategy was dendronized polymer material, bearing as side chains the alternating Fréchet type poly(aryl ether) and Newkome type poly(aliphatic ester) dendrons of the 3rd generation [75,392]. The synthesis was performed via orthogonal double-click alkyne-azide cycloaddition and Diels-Alder reaction between poly(aryl ether) and poly(aliphatic ester) dendrons containing alkyne and furan-protected maleimide groups at their focal points, respectively, with linear polystyrene carrying azide and anthracene moieties in alternating blocks (Figure 22) [392].

3.2.2. Convergent Dendrimer Grafting to Silica-Based Supports

Rhee and Chung suggested the grafting of ready synthesized PAMAM dendrimers to amorphous silica via the reaction with (3-glycidoxypropyl)trimethoxysilane (GPTMS), followed by post-modification with chiral moieties on periphery (Scheme 41) [92]. The materials obtained exhibited a noticeable negative dendritic effect in the enantioselective addition of diethylzinc to benzaldehyde; conversion and enantiomeric excess did not exceed 80% and 20%, respectively and decreased with the increase in the dendrimer generation [92].
Insertion of spacers between the dendrimer and chiral groups (Scheme 41) resulted in a slight increase in conversion and enantiomeric excess due to the better access of substrates to catalytic centers and weakening of multiple interactions among the chiral groups on the dendrimer surface [92]. The maximum conversions (up to 85%) and enantiomeric excess (up to 37%) were found for the catalysts based on dendrimers of the 3rd and 4th generations. However, such catalysts gradually lose their activity at recycling [92].
The similar approach was applied for immobilization of G3 PAMAM dendrimers on the inner surface of titania nanotubes (TiO2 NT) [73]. It was discovered that anchoring of PAMAM dendrimers enhanced drug loading capacity and drug release properties of TiO2 NT, simultaneously decreasing the cytotoxicity of the latter.
Vallet-Regí et al. offered the immobilization of PPI dendrimers of different generations in the pores of the ordered SBA-15 material, using preliminary dendrimer modification by (3-isocyanatopropyl)triethoxysilane (Scheme 42) [74]. The composites synthesized appeared as prospective drug deliverers, with the drug dosage directly controlled by the dendrimer generation via electrostatic and host–guest interactions [74].
Jayaraman and co-workers developed a heterogeneous Pd complex catalyst, based on a phosphine-terminated PETIM dendrimer of the 1st generation, anchored to amorphous silica, modified with (3-chloropropyl)trimethoxysilane (Scheme 43) [102]. Thus, a synthesized hybrid SiO2-PETIM-PPh2-Pd catalyst appeared highly active in the hydrogenation of various olefins under the ambient conditions (25 °C, 1 atm. of H2, AcOEt). The yields of 94–100% were achieved within 1–5 h. The catalyst activity and reaction rate were strongly affected by the negative dendritic effect and decreased for sterically hindered substrates, especially with the internal C=C double bonds; maximum reaction rates were observed for styrenes with terminal C=C double bonds, while C=O double bonds and phenyl rings in the substrate structure remained untouched [102]. The catalyst SiO2-PETIM-PPh2-Pd demonstrated extremely high stability at reuse and storage, maintaining 100% of styrene during 10 reaction cycles (25 °C, 1 atm. of H2, 1 h, AcOEt) [102]. Moreover, under the reaction conditions at recycling, full reduction of Pd2+ to Pd0 took place, and Pd nanoparticles of 2–4 nm in diameter were formed in situ without any Pd leaching or particle agglomeration observed [102].
Maximov and Karakhanov suggested the grafting of PPI dendrimers of the 3rd generation to silica-polyamine composites (SPC) [395] via the Mannich reaction (Scheme 44) [171]. Such hybrid organo-inorganic PPI dendrimer-containing material was further applied as a carrier for both mono- and bimetallic nanoparticles [171,393]. Two-step impregnation of Pd(OAc)2, followed by reduction in the presence of NaBH4, resulted in well-dispersed Pd nanoparticles with a mean particle size of ~2 nm, stabilized by both anchored PPI dendrimers and poly(allyl amine), simultaneously [171].
The synthesized SPC-PPI-G3-Pd catalyst demonstrated very high activity in the hydrogenation of styrene, phenylacetylene, and conjugated dienes (70 °C, 10–30 atm. of H2, 15–60 min, benzene or toluene as solvent) with TOF values reaching 100,000–230,000 h−1, and thus rather exceeding those for SPC-Pd catalysts without grafted dendrimers in the structure [171]. The separate hydrogenations of styrene and phenylacetylene in the presence of a SPC-PPI-G3-Pd catalyst proceeded with similar rates; nonetheless, by carefully adjusting the reaction conditions (the reaction time, hydrogen pressure, and substrate/Pd ratio) it was possible to selectively convert phenylacetylene to styrene with a yield of ~82%.
The product distribution in the hydrogenation of dienes was significantly dependent on the intermediate stability [288,289] and C=C double bond accessibility in both the substrate and alkenes formed. The total selectivity on alkenes reached 93–99%, even at quantitative and near to quantitative diene conversions and hydrogen pressured of 30 atm. (Scheme 45) [171], which rather exceeded that for Alper’s catalysts, based on PAMAM dendrons applied under much milder ambient conditions (25 °C, 1 atm. of H2, MeOH) [148]. Herein, the products of hydrogen trans-1,4-addition were predominant, in contrast to Pd nanocatalysts based on PAMAM or aryl-cored poly(carbosilane/triazol/aryl ether) dendrimers [148,150]. The SPC-PPI-G3-Pd catalyst proved its recyclability and could be reused several times with a slight decrease in activity due the mechanical losses [171].
Silver addition to palladium, supported on an SPC-PPI-G3 hybrid carrier, by the co-complexation method resulted in a drastic decrease in the catalyst activity and TOF values, not exceeding 30,000 h−1, but also in the noticeable increase in the selectivity on styrene in the phenylacetylene hydrogenation (80 °C, 10–30 atm. of H2, 15–60 min, solventless) [393]. By carefully adjusting the reaction conditions (the reaction time, hydrogen pressure, substrate/Pd and Pd/Ag ratios) it was possible to achieve a styrene yield as high as 89–93%.
Hydrogenation of 1-octyne in the presence of SPC-PPI-G3-Pd/Ag catalysts under the said conditions was accompanied by the isomerization of C=C double bond in 1-octene formed that was typical for Pd-containing catalysts [116,124]. It should be noted that 1-octyne and resultant octenes easily underwent hydrogenation compared with phenylacetylene, and total yield of octenes did not exceed 83%, with the selectivity on 1-octene not exceeding 77% among the other hydrogenation products [393]. Nonetheless, the hybrid bimetallic dendritic SPC-PPI-G3-Pd/Ag catalysts appeared highly stable and were recycled several times without any loss in activity and selectivity.
Another approach for the synthesis of hybrid organo-silica ready dendrimer-containing materials consisted in the co-hydrolysis in situ of Si(OEt)4 with PPI or PAMAM dendrimers, peripherally modified with (3-glycidoxy)propyltrimethoxysilane (Scheme 46) [396]. The silica channels were constructed around the dendrimer templates via the sol-gel-method, resulting in microporous material with a pore size of 1–2 nm (Figure 23a). Co-hydrolysis in the presence of additional polymer templates, such as Pluronic P123, favored the formation of SBA-15 type mesoporous material with pore diameters of 5.5–7.5 nm (Scheme 47, Figure 23b) [396]. In the presence of the polymer template, but in the absence of Si(OEt)4, it was possible to obtain a mesoporous dendrimer network with the silanol groups in the node’s lack of the silica matrix (Scheme 48) [193]. It should be noted that, previously, the sol-gel-method via co-hydrolysis in situ was developed for organo-silica materials of SBA-15 and MCM-41 types, containing aromatic [397,398], bipyridyl [399] and chiral tartamide moieties in the matrix structure [400].
Thus, synthesized organo-silica dendrimer-based materials were applied as carriers for Pd and Ru nanoparticles, whose mean size and distribution appeared to be significantly dependent on the carrier structure and dendrimer generation [193,396]. Herein, Ru nanoparticles mostly formed as dendrimer-encapsulated, with the mean particle sizes of 1.0 and 1.3 nm for SiO2-PPI-G3-Ru and meso-SiO2-PPI-G2-Ru, respectively (Figure 24) [396], whereas Pd nanoparticles preferentially formed as dendrimer-stabilized with the mean particle sizes of 2.7–3.3 nm (Figure 25) [193]. Moreover, the shape of particle size distribution for the microporous catalyst SiO2-PPI-G3-Pd appeared similar to the pore size distribution of the initial carrier (Figure 25b) [193].
Dendrimer-based organo-silica Pd-containing catalysts were studied in the hydrogenation of various styrenes, alkynes, and dienes (80 °C, 10–30 atm. of H2, 15–60 min) [193]. In a whole, the catalytic behavior of meso-SiO2-PPI-G2-Pd and SiO2-PPI-G3-Pd was very similar to that for SPC-PPI-G3-Pd [171], and by carefully adjusting the reaction conditions (the reaction time, hydrogen pressure and substrate/Pd ratio) it was possible to achieve total alkene yields of 83–98% in the hydrogenation of alkynes and dienes [193]. Herein, the mesoporous meso-SiO2-PPI-G2-Pd catalyst appeared to be the maximum activity in the hydrogenation of styrenes, including bulky and rigid trans-stilbene, 4-phenylstyrene, and 1,1-diphenylethylene—in spite of clear-cut negative dendritic effect (Figure 26)—with TOF values exceeding 231,000 h−1 for styrene [193], which was similar to those reached in the presence of the SPC-PPI-G3-Pd catalyst [171].
The highest conversions and total alkene selectivities of up to 99% in the hydrogenation of terminal alkynes, tiny and flexible conjugated dienes, as well as of +I-substituted 4-tert-butylstyrene, were found for the microporous catalyst SiO2-PPI-G3-Pd, based on dendrimers of the 3rd generation and, therefore, providing a larger amount of donor amino groups, increasing electron density on Pd nanoparticles and thus facilitating the desorption of alkene formed [128,171,193,285]. Hence, the competition between steric and electron factors took place, and maximum TOF values reaching up to 400,000 h−1 (Figure 27) were achieved for 1-hexyne, isoprene, and 2,4-dimethyl-2,4-hexadiene [193]. The catalyst meso-PPI-G2-Si-Pd, obtained without silica matrix formation, was significantly inferior in performance both to meso-SiO2-PPI-G2-Pd and meso-SiO2-PPI-G2-Pd; nonetheless, it was the only catalyst providing the trans-cyclooctene formation with selectivity of 90–95% [193].
Ru nanoparticles containing catalysts based on PPI dendrimers immobilized in silica pores by co-hydrolysis in situ were studied in the hydrogenation of alkylphenols, levulinic acid, and its esters in water medium [396,401]. The main reaction products of phenol hydrogenation (70–85 °C, 10–30 atm. of H2, H2O, 0.5–6 h, water) were corresponding cyclohexanols with total yield up to 98% [396]. Herein, cis-alkylcyclohexanols were predominant (up to 70% selectivity) at less reaction times (0.5–2 h), especially for bulky and/or ortho-substituted phenols (e.g., o-allylphenol or p-tert-butylphenol), while prolongation of reaction time (up to 6 h) resulted in the increased portion of trans-alkylcyclohexanols (up to 90%) due to the impact of basic dendrimer amino groups, stabilizing the intermediate enol form and thus facilitating its rotation [396,402,403]. A hydrogenolysis, the typical feature of Ru-based catalysts [1], was observed; its portion reached 20–30% among the reaction products, gradually decreasing from cresols to p-tert-butylphenol [396].
Hydrogenation of levulinic acid and its esters mostly proceeded through the preferential cyclization to intermediate angelica-lactones, followed by reduction to γ-valerolactone (GVL) (Scheme 49), which was the major and desired reaction product (80–120 °C, 30 atm. of H2, H2O, 2–6 h, water) [401]. Basic PPI dendrimer amino groups, being able to deprotonate the levulinic acid molecules or to catalyze levulinic ester hydrolysis, were assumed to be responsible for the said reaction pathway (Scheme 50). Hence, for the SiO2-PPI-G3-Ru catalyst, based on dendrimers of the 3rd generation with the larger amount of basic amino groups, the selectivity on γ-valerolactone exceeded 90% even at conversions of 10–50% [401].
For hydrogenation of both phenols and levulinic acid and its esters, the strong negative dendritic effect and competition between steric and electron factors took place, and mesoporous and a sterically less hindered meso-SiO2-PPI-G2-Ru catalyst appeared much more active than a microporous, sterically more hindered SiO2-PPI-G3-Ru catalyst, reaching TOF values up to 6000 h−1 versus 1800 h−1 for phenols [396] and 1830 h−1 versus 930 h−1 for levulinic acid and its esters [401]. The increase in the substrate size resulted, on the one hand, in the respective decrease in activity for both catalysts, and on the other hand, it was surpassed by the sufficient +I-effect of substituent in the substrate, such as in o-allylphenol, p-tert-butylphenol, or ethyl levulinate, resulting in the increased activity for both catalysts (Figure 28) [396,401].
Dendrimer-based organo-silica was easily recoverable and proved their recyclability in the repeated hydrogenation of various substrates (styrene, 1,3-cyclohexadiene, and 2,5-dimethyl-2,4-hexadiene for Pd catalysts [193]; phenol and levulinic acid for Ru catalysts [396,401]) with just a slight decrease in activity due to the mechanical losses.

4. Dendrimer Cross-Linking

Dendrimer cross-linking, resulting in three-dimensional network formation, is one of the simplest approaches for the synthesis of heterogeneous dendrimer-based catalysts, not requiring the special carrier. It can be performed in two different ways. The first is the complex formation with transition metal ions, which are able to bind dendrimer end groups and thus provide the one-step synthesis of the recoverable network dendritic catalysts presenting as coordination polymers [124]. The second way implies the covalent dendrimer cross-linking, using various bi- or polyfunctional agents (diepoxides, diisocyanates, polycarboxylic anhydrides, etc.) [241]. Both of these methods are further considered in the present review.

4.1. Coordination Dendritic Network Polymers

In 2009, Maximov and Karakhanov reported the in situ formation of the heterogeneous dendritic coordination network during the Wacker oxidation of linear C6–C12 alkenes in the presence of CN-terminated PPI dendrimers and 2,2′-biquinoline-4,4’-dicarboxylic acid N,N′-bis(3,3′-iminodipropionitrile)diamide (BQC-IPN) (PdSO4/CuSO4, 80 °C, 5 atm. of O2, H2O/EtOH, 1 h) [124,241]. It was established that, initially, the homogeneous catalytic system precipitated quickly in the beginning of the reaction, forming agglomerates of 50–200 nm in size, containing small particles of 1.5–2 nm, which were attributed to contracted dendrimer molecules (Figure 29) [124].
It was assumed that Pd2+ ions mostly coordinated on the dendrimer terminal C≡N groups, whereas Cu2+ ions did so on the internal tertiary amino groups or bipyridyl moieties, resulting in the sterically hindered dendritic network acting as nanoreactors for the selective oxidation of terminal alkenes (Figure 30) [124,241]. An increase in the dendrimer generation and, therefore, in the CN/Pd ratio, led to the decrease in the reaction rate (negative dendritic effect), but favored the selective methylketone formation, reaching 94% and 90% at 1-octene conversions of 37% and 27% for catalysts PdII/CuII/DAB(CN)16 and PdII/CuII/BQC-IPN, respectively [241]. Moreover, the use of simple, low-molecular nitrile ligands (CH3CN or PhCN) instead of high molecular regular branched dendrimers at the same CN/Pd ratio only retarded the reaction, but did not suppres the oxidation of isomeric alkenes, and inevitably formed in the presence of Pd catalysts [124,241].
Reetz and Giebel developed heterogeneous Lewis acid catalysts based on triflate-terminated PPI dendrimers of the 4th generation, cross-linked with scandium ions (Scheme 51) [404]. The synthesized material revealed the high activity and chemo-selectivity in the Mikayama, Diels-Alder, and Friedel-Crafts reactions, giving yields of target product up to 90% (Scheme 52), which authors explained in terms of the dendritic matrix swelling in water and, as a consequence, the facilitated substrate access to the inner catalytic centers [404]. The catalyst appeared easily recoverable and maintained its efficacy in several repeated cycles.
Majoral and Caminade suggested the coordination network catalyst, based on phosphazene-cored arylphosphite dendrimers, modified by 15-membered triolefinic triaza-macrocycles on periphery (Scheme 53) [405]. Neutral Pd species and clusters originated from Pd2(dba)4 (where dba was dibenzylidene acetone) or Pd(PPh3)4 coordinated on C=C double bonds of macrocycles on the dendrimer periphery, thus bonding neighboring dendrimers (Figure 31). Additionally, S=P arising dendrons could form nanoparticle-cored micelles (Figure 31). Hence, there were ensembles obtained typical for other coordination dendrimer-based materials (Figure 32) [124,405].
Depending on the synthesis conditions and certain structure of dendrimer used, both metal complex and nanoparticle (2–5 nm) catalysts were obtained, which were further tested in the Heck cross-coupling reaction between iodobenzene and butyl acrylate (Bu3N, THF, 60 °C, 24 h) [405]. In spite of the fact that all catalyst samples were initially isolated and could be recovered as solids, due to the low binding density, most of them behaved as homogeneous catalysts in the process.
The best results with the yield of butyl cinnamate up to 96% were achieved for homogeneous catalysts: metal complex and Pd7 cluster. Also, these catalysts appeared easily recoverable and were recycled five times without any loss in activity [405]. This might be due to the well retention of Pd (0) species and small Pd7 clusters by the dendrimer-appended macrocycle moieties, and suitable in size. The heterogeneous catalyst was inferior in activity to the homogeneous metal complex and cluster catalysts, giving a yield of butyl cinnamate not exceeding 89%.
A strong negative dendritic effect was revealed: increase in the dendrimer generation from 0th to 4th, as a consequence in the size of Pd particles, resulted in the crucial decrease in conversion and subsequent Pd leaching at catalyst recycling [405]. Pd nanoparticles, located in the inner dendrimer cavities, appeared hardly accessible for substrate due to the multiple steric hindrances arising from the structure of 4th generation dendrimer [405]. Moreover, in the case of aryl phosphite dendrimers, there was possible migration of nanoparticles from the dendrimer inner cavities to the outer surface, and, further, to the reaction medium [406].

4.2. Covalent Cross-Linked Dendrimer Networks

Crooks and Bergbreiter, for the first time, used networks based on PPI or PAMAM dendrimers cross-linked with poly(3-methylthiophene) or poly(maleic anhydride)-alt-poly(methyl vinylether) copolymer for developing thin films (4–43 nm), appearing to be thermo- and pH-responsive and have electroconductive and selective permeability properties, depending on the treatment conditions [314,407].
Independently, Maximov and Karakhanov suggested the cross-linking of PPI and PAMAM dendrimers with diisocyanates of various size and rigidity, and they were the first who offered to apply the obtained cross-linked dendrimer networks as carriers for heterogeneous nanocatalysts (Scheme 54, Figure 33) [241]. Herein, amino, amido, and/or hydroxyl groups in the structure of dendritic network carriers were responsible for the effective binding of metal ions and, as a consequence, for immobilization of metal nanoparticles within a three-dimensional polymeric matrix consisting of the dendrimer moieties in the nodes, bridged with diurethane or aminoethanol ribs. The properties of the resultant materials, such as mean particle size, catalyst activity, and selectivity, were strongly influenced by the dendrimer nature and generation, as well as by the size, rigidity, and polarity of cross-linking agents, and even by the metal deposition conditions [122,127,128,241,287,408,409], thus making dendrimer network carriers partly similar to the metal-organic frameworks (MOF), whose cell parameters and, therefore, the absorption, catalytic, and transport properties were analogously dependent on both size, nature, and rigidity of the organic ribs and in external conditions (temperature, solvent, pressure) simultaneously [410,411].
Thus, PPI and PAMAM dendrimer network-based catalysts were developed, containing Rh, Ru, and Pd nanoparticles successfully applied for hydrogenation of phenols [122,127,412], aromatic [127,408,413,414], and unsaturated compounds under the moderate conditions (50–90 °C, 10–30 atm. of H2) [123,128,241,287,409,415]. The main reaction products were the same as for dendrimer-based silica supported catalysts [193,396], and typical for the metal nature under the applied conditions; after hydrogenation of aromatic compounds and phenols in the presence of Rh and Ru containing catalysts, there were corresponding alkylsubstituted cyclohexanes, cyclohexanols, and dihydroxycyclohexanes [1,122,127,408,413,414]. In the presence of Pd-containing catalysts, based on the dendrimer networks, styrenes and dienes were selectively converted to corresponding ethylbenzenes and monoenes with the aromatic ring or conjugated C=C or C=O double bonds remained untouched [123,128,241,287,409,415].
In particular, it was found out that Ru catalysts, based on PPI dendrimers, cross-linked with hexamethylene-1,6-diisocyanate (HMDI) [122], were superior to those of Rh, based on PAMAM dendrimers, analogously cross-linked with 1,6-hexamethylene diisocyanate [412], in the hydrogenation of mono- and dihydric phenols, exhibiting TOF values up to 240,000 h−1 (Figure 34) [122]. Analogously, Pd nanoparticles, immobilized in the networks of PPI dendrimers, cross-linked with 4,4′-(3,3′-dimethoxy)diphenyl diisoxyanate (DMDPDI), revealed the efficacy and selectivity in the hydrogenation of phenylacetylene, α,β-unsaturated compounds and conjugated dienes, which were higher than those for Pd catalysts based on DMDPDI cross-linked PAMAM dendrimers (Figure 35) [128,409]. At the same time, PAMAM dendrimer-based catalysts appeared as more active in the hydrogenation of terminal linear alkenes and styrenes (Figure 35) [123,128]. The substrate selectivity for Pd catalysts, based on the networks of both PPI and PAMAM dendrimers, should be remarked, appearing in the highly superior activity for styrenes, dienes, and α,β-unsaturated compounds (i.e., containing the conjugated C=C double bonds) in comparison with linear and cyclic alkenes (1-octene, cyclohexene, etc.) by 15–90 times [128,241].
As for catalysts based on PPI dendrimers and immobilized in the silica pores [396], and those based on the networks of PPI dendrimers exhibited a noticeable competition between steric and electron factors. In most cases, the catalyst activity decreased with the increase in the substrate size and dendrimer generation [122,128,241,408,413] or when passing to the dendrimer network carrier, synthesized with the use of a less expanded linker (e.g., toluene-2,4-diisocyanate (TDI) instead of methylene-bis(p-phenylene diiosocyanate) (MPDI)) (Figure 33) [241]. On the other hand, the presence of strong +I or +M-substituents in the substrate structure [122,408] or the latter closely fitting to the network matrix mesh [409,413] may enhance the catalyst activity even for bulky substrates, such as 4-tert-butylphenol [122], 4-tert-butylstyrene [413], ethyl- and butylbenzenes, etc. [408]. A positive dendritic effect appearing in the increase in the reaction rate and selectivity with the increase in the dendrimer generation was observed in the hydrogenation of styrenes and phenylacetylene in the presence of Pd catalysts based on PPI dendrimers cross-linked with relatively long and flexible linkers, such as hexamethylene diisocyanate or glycerol triglycidyl ether [287,409], which was possibly due to the better stabilization and higher electron saturation of Pd nanoparticles by multiple donor amino groups in the case of PPI dendrimers of the 3rd generation.
The influence of donor amino groups was found to be essential for the phenol hydrogenation in the presence of Rh and Ru catalysts, especially with regards to hydrogenation of dihydric phenols, such as hydroquinone and resorcinol (Figure 34) [122,412]. In the last case, PPI dendrimers additionally stabilized the enol form for the intermediate product 1,3-cyclohexanedione, making it more liable to the further hydrogenation, thus drastically reducing the negative dendritic effect [122]. Analogously, the higher dendrimer content in the network carriers (40–50% vs. 15–20% for organo-silica carriers with the dendrimers, immobilized in the silica pores [396]) resulted in the stabilization and facilitation of the rotation of the adsorbed enol form of intermediate alkyl-substituted cyclohexanone [396,403], thus favoring a higher portion of the corresponding trans hydrogenation product [122]
The use of a polymeric template on the stage of the network carrier synthesis (Scheme 55) along with the bulky and rigid dimethoxydiphenyl diisoxyanate linker afforded to develop highly expanded matrix, favoring to the formation of larger particles (5 nm vs. 1.5–2.5 nm, typical for Pd nanoparticles, encapsulated in the networks of PPI dendrimers [287,409]) and, as a consequence, changing their electron properties [128,287]. Thus, there a highly selective catalyst was obtained for phenyl acetylene hydrogenation, maintaining a high styrene yield (~95–96%) even at quantitative conversions, hydrogen pressure of 30 atm., and prolonged reaction times (Figure 36) [128,287]. The maximum TOF values reached up to 185,000 h−1 [128,287]; moreover, meso-DAB-PPI-G3-DMDPDI-Pd appeared as the only catalyst, maintaining its activity in the phenylacetylene recycling [287]. Another way to change the catalyst electron properties was applied in the development of bimetallic alloyed PdRu nanoparticles, encapsulated in the network of PPI dendrimers of the 3rd generation, cross-linked with hexamethylene diisocyanate, revealing the enhanced activity exclusively in the benzene hydrogenation (Figure 37) [408].
The influence of templates on the catalyst activity can be observed in the example of Ru catalysts, based on the sterically more hindered PPI dendrimer networks [122], which are significantly inferior to those based on PPI dendrimers, immobilized in the silica pores and thus providing better substrate access to Ru nanoparticles [396]. Moreover, sterically hindered dendritic network carriers appeared to be much more sensitive to the substrate shape and geometry, resulting in better activity for para-substituted positively induced phenols, such as p-ethylphenol or p-tert-butylphenol [122,396].
Not only the template, but also the reaction conditions, such as temperature and solvent, on the stage of nanoparticle encapsulation had a great influence on the physical chemical properties and, as a consequence, on the hydrogenation activity, which was demonstrated on the example of Ru catalysts, based on PPI dendrimers cross-linked with poly(ethylene glycol) (PEG) diglycidyl ether [127]. Consisting of the blocks with differing low critical solution temperatures (LCST) (PEG chains and PPI dendrimers), these catalysts revealed different thermos-responsive properties on the hydrogenation of both phenol and benzene, essentially depending on the reaction conditions on the stages of both catalyst synthesis and subsequent hydrogenation. It was found out that the use of the “bad” reaction conditions in the catalyst synthesis (inappropriate, “bad” solvent, e.g., THF instead of water, or temperature above LCST for polymer blocks, constituting the network support) favored the superior activity in the hydrogenation of phenol and, especially, of benzene, but resulted in the partial or complete loss of thermo-responsive properties [127] similar to thin films, based on PPI or PAMAM dendrimers, cross-linked with poly(maleic anhydride) [314].
Deposition of Rh or Pd nanoparticles on the networks of PPI and PAMAM dendrimers under the conditions of supercritical CO2 (scCO2) resulted in the extremely efficient catalysts, exhibiting TOF values up to 280 s−1 (>1,000,000 h−1) and 207 s−1 (>745,000 h−1), respectively, at the metal loadings as low as 0.01–0.29% [416]. Herein, Rh catalysts, based on PAMAM dendrimers, revealed the highest activity in the hydrogenation of alkenes, such as 1-octene and, especially, styrene, and a noticeable negative dendritic effect was observed for catalysts on both PAMAM and PPI dendrimers. Pd catalysts appeared to be the most effective in the hydrogenation of conjugated dienes [415].
The presence of water was found to be crucial for hydrogenation of phenols and aromatic compounds in the presence of Ru catalysts [122,408,413]. Simultaneously, the dendritic matrix, cross-linked with diisocyanates, was subjected to Ru-catalyzed hydrolysis [122]. Nonetheless, the catalysts, based on the dendritic networks, can be successfully reused several times without a significant loss in activity and selectivity under both aqueous (Ru, Rh) and non-aqueous conditions (Pd) [122,128,287].
Moreover, the network matrices, based on PPI dendrimers, appeared to be a suitable carrier for Ni-W sulfide catalysts, and successfully applied for the hydrogenation of naphthalene (350–400 °C, 50 atm. of H2, 5 h) [414]. The catalyst efficiency was found to be dependent on the synthesis conditions (ex situ or in situ), process temperature, and sulfidizing agent, and the best results (55% selectivity on decalins and 45% selectivity on tetralin at 98% conversion of naphthalene) were reached at 380 °C using the catalyst synthesized in situ and elemental sulfur as the sulfidizing agent [414]. It should be noted that PPI dendrimers may undergo retro-Michael addition under hydrogenation conditions even at temperatures below 100 °C [61]; as a consequence, dendrimer network carriers, at least, were partially decomposed at 350–400 °C and 50 atm. of H2, under which the process was performed [394]. This might be an additional reason as to why in situ catalysts appeared more efficient.
Ogasawara and Kato developed a mesoporous polymer Tailor-made catalyst based on PAMAM dendrimer-encapsulated nanoparticles [417]. This approach suggested the formation of small dendrimer-encapsulated Pd nanoparticles (~2 nm) in situ, simultaneously with the synthesis of the cross-linked polymeric matrix, outside the cavities, formed by PAMAM dendrimers (Scheme 56). The catalyst synthesized revealed high activity in the Suzuki cross-coupling between phenylboronic acid and 4-bromoacetophenone (80 °C, K2CO3, H2O), giving a quantitative conversion within 4 h with TOF values up to 2500 h−1 [417]. Replacement of the –M-acceptor acetyl group by the C≡N group, possessing with the strong coordination effect, or by +M-donor OH or OCH3 groups, resulted in a noticeable decrease in the catalyst activity. The catalyst exhibited good resistance to metal leaching and can be recycled eight times, maintaining the yield of cross-coupling target product of more than 90% [417].
The similar approach was applied by Nabavinia, Kanjilal, and Noshadi for the synthesis of PEI dendrimer-encapsulated palladium nanoparticles, encaged into the resorcinol formaldehyde micro/mesoporous resin polymer network (MPR) in situ in the absence of the polymer template (Scheme 57) [416]. The catalyst activity and recyclability in the Suzuki cross-coupling reaction between phenylboronic acid and toluilhalide was found to be strongly dependent on the conditions of both catalyst synthesis (the solvent used) and cross-coupling reaction (arylhalide nature, temperature, solvent). The maximum efficacy was observed for the catalyst synthesized in water-ethanol medium, and the quantitative conversions for 4-iodotoluene were observed already within 30–50 min at 85 °C using water-ethanol medium again and K2CO3 as a base. For 4-bromotoluene, the conversion of about 80% was reached within 5–6 h at 110 °C using water-DMF medium and K2CO3 as a base. The synthesized PEI-MPR-Pd catalysts revealed high thermal stability and good resistance to metal leaching and thus can be recycled several times without a noticeable loss in activity and used in the continuous-flow micro set-up, with conversion reaching 62% at 85 °C and a flow rate of 0.2 μL/min [416].

5. Magnetically Separable Heterogeneous Dendrimer-Based Catalysts

The use of nanoparticles and carriers possessing ferromagnetic properties, opens the way for versatile design of the easily and effectively recoverable catalysts, essentially diminishing the metal losses and, therefore, the cost of the process and target products [10,62,64,418,419]. Moreover, the presence of iron or nickel oxide in the nanoparticle core provides not only the easy catalyst separation, but also is able to enhance the catalytic activity selectivity, and even to change the reaction pathway [418] that has been demonstrated on the examples of Ru-catalyzed nitrobenzene hydrogenation [420], Zn-catalyzed methanol synthesis from syngas [421], selective Ru-catalyzed hydrogenolysis of cellulose to ethylene glycol in water [422], and zeolite ZSM-5 catalyzed selective methanol conversion to C5–C8 and C9–C11 hydrocarbons (MTH) [423].
With respect to dendrimer-based materials, magnetic separation can be applied to both homogeneous and heterogeneous catalysts [10,62,64,419]. In the first case, it is realized through the dendrimer-encapsulated mono- or bimetallic core-shell nanoparticles, where the core is magnetically active metal or metal oxide (Fe, Co, Ni) and the shell is a catalytically active metal (Ru, Pt, Pd, Rh etc.) (Figure 38, left); on the other side, there can be a synthesized nanoparticle-cored dendrimer, where there is magnetically active metal or metal oxide again, while catalytic centeres are located elsewhere in the dendron moiety (Figure 38, right) [419,424,425]. Thus, the iron oxide (Fe2O3) nanoparticle cored melamine dendrimer, functionalized with Pd phosphine complexes on the termini (Figure 39), proved its efficacy in the Suzuki cross-coupling reaction between various arylhalide and meta-substituted arylboronic acids (K2CO3, CH2Cl2/DMF, 12 h), giving yields of 72–81%, depending on the halide and substituent nature [425]. In spite of the occurrence as homogeneous catalyst, this hybrid material appeared easily recoverable under the external magnetic field and was used in the subsequent reaction cycles without any loss in activity [425].
Heterogeneous dendrimer-based magnetically separable catalysts can be synthesized in different ways [419]. Nonetheless, in the structure of such catalysts there can be distinguished the following main constituents (Figure 39) [419,425]: (1) the magnetic core (Fe2O3, Fe3O4, Co, Ni), which may be additionally coated with silica or polymer; (2) dendrimers or dendrons, immobilized on the magnetic core using covalent attachment (convergent or divergent), π–π interactions, or hydrogen bonds; (3) catalytic centers (metal complexes or nanoparticles), located inside or outside the dendrimer moiety. Several examples for the synthesis and application of magnetically separable dendritic catalysts are further considered in the present review.

5.1. Ferrous Oxide Core Magnetically Separable Dendritic Catalysts

Bronstein et al. developed hybrid magnetically sensitive material Den-Fe3O4 based on Müllen-type polyphenylene pyridyl-terminanated dendrons of the 1st generation, covalently attached to iron oxide nanoparticles of 18–25 nm in size (Scheme 58) [426,427,428]. Subsequent impregnation with Pd2+ ions, coordination on pyridyl moieties, followed by molecular hydrogen reduction, resulted in well dispersed Pd nanoparticles of 0.9–1.9 nm in diameter, uniformly distributed on the surface of iron oxide (Figure 40) [426,427,428,429]. Herein Pd species functioned as peculiar linking nods between dendrimer-coated iron oxide nanoparticles—similar to Pd2+, Cu2+ and Sc3+ ions in the dendrimer coordination network catalysts [124,241,404]—thus facilitating the catalyst aggregation and separation under the magnetic field [426].
The synthesized catalyst Pd-Den-Fe3O4 proved its efficacy in the Suzuki cross-coupling reaction between phenylboronic acid and various 4-bromoarenes [426,427] and semi-hydrogenation of dimethylethynylcarbinol (DMEC) to dimethylvinylcarbinol (DMVC) [428,429]. In the last case, the yields of target product and TOF values as high as 90–98% and 5760–33,480 h−1, respectively, were reached already after 0.5–1 h at 90 °C and atmospheric hydrogen pressure (Scheme 59a) [428,429]. Moreover, tha Pd-Den-Fe3O4 catalyst maintained its activity and highest selectivity on dimethylvinylcarbinol in several repeated cycles [428,429].
The best yields as high as 90% in the Suzuki cross-coupling reaction between phenylboronic acid and 4-bromoanisole were reached after 1 h at 70 °C, using K2CO3 as a base and dioxane/water mixture as a reaction medium [426]. The use of hyperbranched pyridylphenylene polymer (PPP), cross-linked with diphenyl ether moieties and Pd(OAc)2, supported on magnetic silica (MS), allowed it to reach quantitative yields in Suzuki cross-coupling with electron withdrawing substrates (4-bromobenzaldehyde, 4-bromonitrobenzene) already within 3–15 min with TOF values up to 23,500 h−1 (H2O/EtOH, Na2CO3, 60 °C) [427].
Moreover, the catalyst Pd-Den-Fe3O4 appeared suitable for the process proceeding in the continuous-flow microreactor where catalyst beds were immobilized, using the constant magnetic field (Scheme 59b) [426]. The best results (conversion of 17–18% and selectivity of 53–55% on 4-phenylanisole) were achieved at 70 °C, flow rate of 0.2 mL/min, and pressure of 5 atm. for double-sided glass reactor and at 90 °C, flow rate of 0.2 mL/min, and pressure of 10 atm. for double-sided stainless-steel reactor, respectively. Herein, no Pd leaching was observed in the eluate after storing the samples for 1 week [426].
Rosario-Amorin, Nlate, and Heuzé suggested the Pd complex catalysts based on bis(aminomethyl-P,P-dialkylphosphine) terminated poly(alkyl aryl ether) dendrons covalently grafted to γ-Fe2O3 nanoparticles of 300 nm in diameter, coated with polycarboxylic acid (Figure 41) [430]. These catalysts revealed a noticeable positive dendritic effect in the Suzuki cross-coupling reaction between arylbromides and phenylboronic acid at 65 °C, using NaOH a base and THF/triton X405/water as a reaction medium (Scheme 60). The maximum yields, as high as 100% within 0.5–1 h, were achieved for arylbromides with strong −M-para-substituents (-C(=O)H, -NO2), whereas +I-substituents, especially in ortho-position to Br or B(OH)2 group in the substrate, vice versa, retarded the reaction. Nonetheless, all of the synthesized hydride dendritic catalysts appeared easily recoverable by means of magnetic separation and successfully reused in 25 reaction cycles with just a slight decrease in conversion (100% → 95%) [430].
Astruc and co-workers have designed well-recoverable materials based on PEG-terminated poly(carbosilane/triazol/aryl ether) dendrimers and dendrons, both covalently attached (Figure 42) and immobilized by means of hydrogen bonds on the surface of silica coating Fe3O4 or Fe2O3 nanoparticles (Scheme 61), called superparamagnetic iron oxide nanoparticles (SPIONs) [10,64,431,432,433]. Impregnated with Pd nanoparticles with a mean size of 1.5–3 nm, these catalysts appeared very effective in the various cross-coupling and alcohol oxidation reactions under moderate conditions (Scheme 62), giving the quantitative yields after 24 h at 60–100 °C and TOF values up to 920 h−1 [432,433]. A strong positive dendritic effect on the catalyst activity and stability was revealed, making possible a successful recovery under the magnetic field and subsequent repeated use up to 10 times [432,433], thus significantly surpassing nondendritic magnetically separable catalysts with γ-Fe2O3@SiO2 core [431]. Herein, bulk dendritic fragments afforded the localization of Pd nanoparticles near the SPION core surface, thus making the catalyst more robust and stable [431].
Alper at al. suggested well recoverable Rh-containing catalysts, based on PAMAM dendrons covalently attached to silica coated iron oxide Fe3O4 (Figure 43) using a divergent strategy for styrene hydroformylation [419,434]. The quantitative conversions and branched/linear ratios of more than 40:1 after 16–20 h were reached at 40–50 °C, 70 atm. of CO/H2 (1:1), using CH2Cl2 as a reaction medium, and the highest branched product yield of up to 99% were obtained for the substrates with −M substituents (e.g., 4-vinylbenzoic acid). An increase in the dendron generation influenced slightly negatively on the branched/linear ratio (44:1 for G1 and 40:1 for G3), but rather positively on the catalyst stability, and conversions of 98–100% maintained during 5 reaction cycles [434].
Os (VI) heterogenized metal complex catalysts, based on the tetraalkylammonium salts, covalently attached to silica coated iron oxide Fe3O4 via (3-aminopropyl)tetraoxysilane bridge and additionally stabilized by poly(aryl ether) dendrons, where the latter one of the substituents at quaternary nitrogens (Figure 44) were successfully applied in the dihydroxylation of both linear and cyclic olefins (Scheme 63) [435]. The diol yields of 87–97% were achieved after 2–9 h under mild conditions (methylmorpholine N-oxide (NMO) as an oxidant, room temperature, acetone/water as a reaction medium), thus being significantly superior to the control catalyst, giving a yield of 85% after 5 h (Figure 44). The positive dendritic effect on both the reaction rate and catalyst stability was found, and the catalyst, based on the dendron of the 2nd generation, proves its reusability in the oxidation of various alkenes, maintaining the diol yields of 90–97% during 5 cycles at least [435].

5.2. Co Core Magnetically Separable Dendritic Catalysts

Caminade, Majoral, and Ouali developed magnetically separable Pd complex catalysts, formed in situ by the interaction of palladium (II) acetate with diphenylphosphite-terminated polyphosphazene dendrimers [436]. The catalyst heterogenization was accomplished using pyrene moiety at the dendron focal point via π–π interactions with graphene-coated cobalt magnetic nanoparticle (Co MNP) (Scheme 64) [419]. Thus, synthesized catalysts were successfully applied in the Suzuki cross-coupling reaction (Scheme 65). The quantitative conversions were reached for −M (-C(=O)CH3, (-C(=O)OH) para-substituted arylbromide, coupled with unsubstituted phenylboronic acid, after 15 h at 60 °C, using Na2CO3 as a base, H2O/THF mixture as a reaction medium, and zero generation dendrons as a stabilizing ligand [436]. During the reaction, the catalyst immobilized and turned to a homogeneous form and was easily recovered under the external magnetic field by the temperature cooling down to 20–25 °C, thus resulting in the π–π interactions renewal [419,436]. Hence, using this release and catch strategy, the catalyst can be successfully reused up to 12 times without any loss in activity [419,436].

6. Conclusions

Dendrimers are a unique class of organic compounds. Due to their branched, regular structure, and a variety of types and possibilities for modification, dendrimers have found wide applications in nanomedicine, chemical engineering, light and electron sensing devices, and, especially, in nanocatalysis. The present review has compiled the common methods for the synthesis and application of the dendritic catalysts, especially emphasizing the advances in design of the heterogeneous dendrimer-based materials during the last two decades. As one can overview, in spite of the variety of the synthetic routes, developed for heterogeneous dendritic catalyst, most of them can be narrowed down to three main approaches.
The simplest is the wetness impregnation of various heterogeneous carriers, such as ordered mesoporous silica, carbon black, carbon nanotubes, or graphite with the colloidal solution of dendrimer-encapsulated nanoparticles. The resistance to metal leaching and, therefore, the catalyst recyclability are provided, and on the one hand, by hydrogen bonds or π–π interactions between dendrimer molecules and carrier, and, on the other hand, by the retention of metal nanoparticles inside the cavities of higher generation dendrimers—similar to homogeneous dendritic catalysts. Thus, synthesized catalysts appeared suitable for the severe industrial conditions and processes, such as hydrocarbon hydrogenation, dehydrogenation, ring-opening, isomerization, oxidation, etc.
The second approach is the covalent grafting of dendrimers or dendrons to heterogeneous carriers (amorphous and mesoporous silica, polystyrene, carbon nanotubes), thus stabilizing the dendritic ligand. This strategy can be realized both in convergent and divergent ways; the convergent way, using ready-prepared dendrimers and dendrons, seems to be advantageous, enhancing the coating yield and diminishing the defects in the structure of future material. Herein, the certain synthetic route strongly depends on both the dendrimer and carrier nature. Dendrimer anchoring to silica-based supports results in the hybrid organic-inorganic, or organo-silica materials.
The third approach implies the synthesis of dendrimer networks carriers. Dendrimers can be bound here by transition metal ions, such as Pd, Cu, Sc, etc., forming coordination polymer networks, acting as acid-bases and metal complex catalysts. Behavior of such catalysts is strongly dependent on the dendrimer nature. Thus, sterically hindered PPI dendrimer of higher generations (3rd and 4th) are able to provide the best reaction yield and selectivity. Vice versa, phosphazene-cored aryl phosphite dendrimers are much more effective at zeroth generation. Otherwise, dendrimers can be cross-linked with diepoxides, diisocyanates, etc., thus resulting in fully organic heterogeneous carriers, which can be speculated as a peculiar analogue of metal-organic frameworks. The properties of these materials and their behavior in catalysis are strongly dependent on the dendrimer nature and generation, on the one side, and on linker size, rigidity, and polarity, on the other side.
Finally, magnetically separable materials can be synthesized, using both weak interactions (hydrogen bonds, π–π interactions) and covalent grafting between dendrimer or dendron and magnetic cores, often coated with silica, carbon, or polymer shell.
Metal appending for both grafted dendritic materials and dendrimer networks is conducted similar to procedures for the synthesis of homogeneous dendrimer-based catalysts, and complex formation between the dendrimer functional node and/or end groups and metal ions or low-molecular complexes is the driving force for the metal impregnation inside the dendritic carrier. Both monometallic and bi- or polymetallic heterogeneous dendritic catalysts can be synthesized in this way. Among others, polyamine dendrimers (PAMAM, PPI, PETIM, PEI) are characterized by commercial ability, relatively simple synthetic procedure and, first of all, the large amount of donor groups able to catch metal and complex formation; hence, the overwhelming majority of the catalyst examples, illustrated in the review, refers to dendrimers of the said types.
Heterogenized dendrimer-based catalysts are obtained and stored as solids/powders similar to conventional heterogeneous catalysts. Nonetheless, various examples in hydrogenation, hydroformylation, oxidation, cross-coupling reactions, etc. have revealed them not only to maintain their efficacy at recycling, being easily recoverable, but also to exhibit performance, often superior to analogous homogeneous catalysts, with TOF values reaching up to 400,000 h−1. Herein, dendrimer moieties acted as nanoreactors, preventing nanoparticle agglomeration and metal leaching and providing a high reaction rate and selectivity. Reaction in these nanoreactors may proceed in a pseudo-homogeneous way on small nanoparticles and clusters (nanoheterogeneous catalysis) or on the metal complex centers, attached to heterogenized dendrimers or dendrons. The “Release and catch mechanism, suggesting the temporary process transition to homogeneous phase, can also take place. Moreover, polyamine dendrimers (PPI, PAMAM, PETIM) in water medium may act as acid-base catalysts, thus additionally increasing the reaction rate and directing the reaction pathway.
Depending on both dendrimer and reaction type, as well as on the substrate electronic properties, together influencing the stability and accessibility of catalytically active species, a positive or negative dendritic effect can take place. In some cases, e.g., semi-hydrogenation of alkynes or conjugated dienes or hydrogenation of phenols, the selectivity can be affected by both the dendrimer type—similar to homogeneous dendrimer-based catalysts—and the dendrimer portion in the carrier.
The catalysts, based on the dendrimer networks, should be especially remarked, as their properties (mean particle size and distribution, thermoresponsivity, efficacy and selectivity) have appeared to be directly influenced by several factors simultaneously. Those are the dendrimer type and generation; linker size, rigidity, and polarity; reaction conditions for the network synthesis and metal deposition (temperature below or above LCST, “good” or “bad” solvent, the presence or absence of polymeric template, supercritical medium etc.). Varying these parameters, one can obtain the efficient catalysts for hydrogenation of linear alkenes to alkanes or, vice versa, of phenylacetylene to styrene—even at quantitative substrate conversions.
Hence, heterogeneous dendrimer-based catalysts have proved their efficacy and recyclability in various chemical processes. In spite of many advances in this field, the design of heterogeneous dendritic catalysts is still developing, featuring versatile approaches.

Author Contributions

Original idea and conceptualization, reviewing by A.M.; writing of the original draft, reviewing, and editing by A.Z. and E.K. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was written as a part of the State Program of TIPS RAS and the Russian Science Foundation grant no. 17-73-30046.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nishimura, S. Handbook of Heterogeneous Catalytic Hydrogenation for Organic Synthesis; John Wiley & Sons, Inc.: New York, NY, USA, 2001; p. 737. [Google Scholar]
  2. Weissermel, K.; Arpe, H.-J. Industrial Organic Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 1993; p. 457. [Google Scholar]
  3. Ertl, G.; Knözinger, H.; Weitkamp, J.; Schüt, F. Handbook of Heterogeneous Catalysis, 2nd ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2008; p. 4270. [Google Scholar]
  4. Augustine, R.L. Heterogeneous Catalysis for the Synthetic Chemist; Marcel Dekker, Inc.: New York, NY, USA, 1996; p. 647. [Google Scholar]
  5. Cornils, B.; Herrmann, W. Applied Homogeneous Catalysis with Organometallic Compounds; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2002; p. 1450. [Google Scholar]
  6. Bellussi, G.; Bohnet, M.; Bus, J.; Drauz, K.; Greim, H.; Jackel, K.-P.; Karst, U.; Kleemann, A.; Kreysa, G.; Laird, T.; et al. Ullmann’s Encyclopedia of Industrial Chemistry, 7th ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2011; p. 29456. [Google Scholar] [CrossRef]
  7. Richards, R. Surface and Nanomolecular Catalysis; CRC Press, Taylor & Francis Group: Boca Raton, FL, USA, 2006; p. 531. [Google Scholar]
  8. Antipin, I.S.; Alfimov, M.V.; Arslanov, V.V.; Burilov, V.A.; Vatsadze, S.Z.; Voloshin, Y.Z.; Volcho, K.P.; Gorbatchuk, V.V.; Gorbunova, Y.G.; Gromov, S.P.; et al. Functional supramolecular systems: Design and applications. Russ. Chem. Rev. 2021, 90, 895–1107. [Google Scholar] [CrossRef]
  9. Centi, G.; Perathoner, S. Creating and mastering nano-objects to design advanced catalytic materials. Coord. Chem. Rev. 2011, 255, 1480–1498. [Google Scholar] [CrossRef]
  10. Yamamoto, K.; Imaoka, T.; Tanabe, M.; Kambe, T. New horizon of nanoparticle and cluster catalysis with dendrimers. Chem. Rev. 2020, 120, 1397–1437. [Google Scholar] [CrossRef] [PubMed]
  11. Cushing, B.L.; Kolesnichenko, V.L.; O’Connor, C.J. Recent advances in the liquid-phase syntheses of inorganic nanoparticles. Chem. Rev. 2004, 104, 3893–3946. [Google Scholar] [CrossRef] [PubMed]
  12. Darr, J.A.; Zhang, J.; Makwana, N.M.; Weng, X. Continuous Hydrothermal synthesis of inorganic nanoparticles: Applications and future directions. Chem. Rev. 2017, 117, 11125–11238. [Google Scholar] [CrossRef] [Green Version]
  13. Chakraborty, I.; Pradeep, T. Atomically precise clusters of noble metals: Emerging link between atoms and nanoparticles. Chem. Rev. 2017, 117, 8208–8271. [Google Scholar] [CrossRef]
  14. Semagina, N.; Kiwi-Minsker, L. Recent advances in the liquid-phase synthesis of metal nanostructures with controlled shape and size for catalysis. Catal. Rev. 2009, 51, 147–217. [Google Scholar] [CrossRef]
  15. Haga, M.; Kobayashi, K.; Terada, K. Fabrication and functions of surface nanomaterials based on multilayered or nanoarrayed assembly of metal complexes. Coord. Chem. Rev. 2007, 251, 2688–2701. [Google Scholar] [CrossRef]
  16. Dawson, R.; Cooper, A.I.; Adams, D.J. Nanoporous organic polymer networks. Progr. Polym. Sci. 2012, 37, 530–563. [Google Scholar] [CrossRef]
  17. Raynal, M.; Ballester, P.; Vidal-Ferrana, A.; van Leeuwen, P.W.N.M. Supramolecular catalysis. Part 1: Non-covalent interactions as a tool for building and modifying homogeneous catalysts. Chem. Soc. Rev. 2014, 43, 1660–1733. [Google Scholar] [CrossRef]
  18. Severinsen, R.J.; Rowlands, G.J.; Plieger, P.G. Coordination cages in catalysis. J. Incl. Phenom. Macrocycl. Chem. 2020, 96, 29–42. [Google Scholar] [CrossRef]
  19. Meeuwissen, J.; Reek, J.N.H. Supramolecular catalysis beyond enzyme mimics. Nat. Chem. 2010, 2, 615–621. [Google Scholar] [CrossRef]
  20. van Dongen, S.F.M.; Elemans, J.A.A.W.; Rowan, A.E.; Nolte, R.J.M. Processive catalysis. Angew. Chem. Int. Ed. 2014, 53, 11420–11428. [Google Scholar] [CrossRef] [Green Version]
  21. Jongkind, L.J.; Caumes, X.; Hartendorp, A.P.T.; Reek, J.N.H. Ligand template strategies for catalyst encapsulation. Acc. Chem. Res. 2018, 51, 2115–2128. [Google Scholar] [CrossRef]
  22. Pappalardo, A.; Puglisi, R.; Trusso Sfrazzetto, G. Catalysis inside supramolecular capsules: Recent developments. Catalysts 2019, 9, 630. [Google Scholar] [CrossRef] [Green Version]
  23. Elemans, J.A.A.W.; Nolte, R.J.M. Porphyrin cage compounds based on glycoluril—From enzyme mimics to functional molecular machines. Chem. Commun. 2019, 55, 9590–9605. [Google Scholar] [CrossRef]
  24. Hasegawa, S.; Tsukuda, T. Exploring novel catalysis using polymer-stabilized metal clusters. Bull. Chem. Soc. Jpn. 2021, 94, 1036–1044. [Google Scholar] [CrossRef]
  25. Sadjadi, S. Chapter 6—Dendrimers as nanoreactors. In Organic Nanoreactors from Molecular to Supramolecular Organic Compounds; Sadjadi, S., Ed.; Academic Press, Elsevier: London, UK, 2016; pp. 159–201. [Google Scholar] [CrossRef]
  26. Karakhanov, E.A.; Maksimov, A.L.; Zolotukhina, A.V.; Kardasheva, Y.S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492. [Google Scholar] [CrossRef]
  27. Gonzalez-Galvez, D.; Lara, P.; Rivada-Wheelaghan, O.; Conejero, S.; Chaudret, B.; Philippot, K.; van Leeuwen, P.W.N.M. NHC-stabilized ruthenium nanoparticles as new catalysts for the hydrogenation of aromatics. Catal. Sci. Technol. 2013, 3, 99–105. [Google Scholar] [CrossRef]
  28. Astruc, D.; Lu, F.; Ruiz Aranzaes, J. Nanoparticles as recyclable catalysts: The frontier between homogeneous and heterogeneous catalysis. Angew. Chem. Int. Ed. 2005, 44, 7852–7872. [Google Scholar] [CrossRef]
  29. Lara, P.; Philippot, K.; Chaudret, B. Organometallic ruthenium nanoparticles: A Comparative study of the influence of the stabilizer on their characteristics and reactivity. ChemCatChem 2013, 5, 28–45. [Google Scholar] [CrossRef]
  30. Hu, H.; Xin, J.H.; Hu, H.; Wang, X.; Miao, D.; Liu, Y. Synthesis and stabilization of metal nanocatalysts for reduction reactions —A review. J. Mater. Chem. A 2015, 3, 11157–11182. [Google Scholar] [CrossRef]
  31. Favier, I.; Teuma, E.; Gómez, M. Palladium and ruthenium nanoparticles: Reactivity and coordination at the metallic surface. C. R. Chim. 2009, 12, 533–545. [Google Scholar] [CrossRef]
  32. Axet, M.R.; Philippot, K. Catalysis with colloidal ruthenium nanoparticles. Chem. Rev. 2020, 120, 1085–1145. [Google Scholar] [CrossRef]
  33. Dupont, J.; Scholten, J.D. On the structural and surface properties of transition-metal nanoparticles in ionic liquids. Chem. Soc. Rev. 2010, 39, 1780–1804. [Google Scholar] [CrossRef]
  34. Maksimov, A.L.; Kuklin, S.N.; Kardasheva, Y.S.; Karakhanov, E.A. Hydrogenation of Phenols in ionic liquids on rhodium nanoparticles. Petrol. Chem. 2013, 53, 157–163. [Google Scholar] [CrossRef]
  35. Kuklin, S.; Maximov, A.; Zolotukhina, A.; Karakhanov, E. New approach for highly selective hydrogenation of phenol to cyclohexanone: Combination of rhodium nanoparticles and cyclodextrins. Catal. Commun. 2016, 73, 63–68. [Google Scholar] [CrossRef]
  36. Nowicki, A.; Zhang, Y.; Léger, B.; Rolland, J.-P.; Bricout, H.; Monflier, E.; Roucoux, A. Supramolecular shuttle and protective agent: A multiple role of methylated cyclodextrins in the chemoselective hydrogenation of benzene derivatives with ruthenium nanoparticles. Chem. Commun. 2006, 3, 296–298. [Google Scholar] [CrossRef] [PubMed]
  37. Noël, S.; Léger, B.; Ponchel, A.; Philippot, K.; Denicourt-Nowicki, A.; Roucoux, A.; Monflier, E. Cyclodextrin-based systems for the stabilization of metallic (0) nanoparticles and their versatile applications in catalysis. Catal. Today 2014, 235, 20–32. [Google Scholar] [CrossRef] [Green Version]
  38. Hapiot, F.; Monflier, E. Unconventional Approaches Involving cyclodextrin-based, self-assembly-driven processes for the conversion of organic substrates in aqueous biphasic catalysis. Catalysts 2017, 7, 173. [Google Scholar] [CrossRef] [Green Version]
  39. Potier, J.; Menuel, S.; Fournier, D.; Fourmentin, S.; Woisel, P.; Monflier, E.; Hapiot, F. Cooperativity in aqueous organometallic catalysis: Contribution of cyclodextrin-substituted polymers. ACS Catal. 2012, 2, 1417–1420. [Google Scholar] [CrossRef]
  40. Tilloy, S.; Binkowski-Machut, C.; Menuel, S.; Bricout, H.; Monflier, E. Phosphane-based cyclodextrins as mass transfer agents and ligands for aqueous organometallic catalysis. Molecules 2012, 17, 13062–13072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Vilas Shinde, V.; Jeong, D.; Jung, S. An amino-chain modified β-cyclodextrin: A supramolecular ligand for Pd(OAc)2 acceleration in Suzuki–Miyaura coupling reactions in water. Catalysts 2019, 9, 111. [Google Scholar] [CrossRef] [Green Version]
  42. Liu, Y.; Hartman, R.L. Reaction kinetics of a water-soluble palladium-β-cyclodextrin catalyst for a Suzuki–Miyaura cross-coupling in continuous flow. React. Chem. Eng. 2019, 4, 1341–1346. [Google Scholar] [CrossRef]
  43. Cocq, A.; Bricout, H.; Djedaïni-Pilard, F.; Tilloy, S.; Monflier, E. Rhodium-Catalyzed aqueous biphasic olefin hydroformylation promoted by amphiphilic cyclodextrins. Catalysts 2020, 10, 56. [Google Scholar] [CrossRef] [Green Version]
  44. Huc, V.; Pelzer, K. A new specifically designed calix[8]arene for the synthesis of functionalized, nanometric and subnanometric Pd, Pt and Ru nanoparticles. J. Colloid Interface Sci. 2008, 318, 1–4. [Google Scholar] [CrossRef]
  45. Zhang, Q.; Catti, L.; Tiefenbacher, K. Catalysis inside the hexameric resorcinarene capsule. Acc. Chem. Res. 2018, 51, 2107–2114. [Google Scholar] [CrossRef] [Green Version]
  46. Gaeta, C.; Talotta, C.; De Rosa, M.; La Manna, P.; Soriente, A.; Neri, P. The hexameric resorcinarene capsule at work: Supramolecular catalysis in confined spaces. Chem. Eur. J. 2019, 25, 4899–4913. [Google Scholar] [CrossRef]
  47. Bekturov, E.A.; Kudaibergenov, S.E. Catalysis by Polymers; WILEY-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2002; p. 153. [Google Scholar] [CrossRef]
  48. Karakhanov, E.A.; Dedov, A.G.; Tskhai, L.E.; Shatalov, V.V. Hydrogenation of oxygen-containing functional derivatives of aromatic hydrocarbons. Petrol. Chem. 1993, 33, 202–227. [Google Scholar]
  49. Zhang, W.; Sun, Y.; Zhang, L. Fabrication of high efficient silver nanoparticle catalyst supported on poly(glycidyl methacrylate)−polyacrylamide. Ind. Eng. Chem. Res. 2016, 55, 12398–12406. [Google Scholar] [CrossRef]
  50. Malakhova, I.; Privar, Y.; Parotkina, Y.; Mironenko, A.; Eliseikina, M.; Balatskiy, D.; Golikov, A.; Bratskaya, S. Rational design of polyamine-based cryogels for metal ion sorption. Molecules 2020, 25, 4801. [Google Scholar] [CrossRef]
  51. Veerakumar, P.; Velayudham, M.; Lu, K.-L.; Rajagopal, S. Silica-supported PEI capped nanopalladium as potential catalyst in Suzuki, Heck and Sonogashira coupling reactions. Appl. Catal. A Gen. 2013, 455, 247–260. [Google Scholar] [CrossRef]
  52. Mori, S.; Ohkubo, T.; Ikawa, T.; Kume, A.; Maegawa, T.; Monguchi, Y.; Sajiki, H. Pd(0)–polyethyleneimine complex as a partial hydrogenation catalyst of alkynes to alkenes. J. Mol. Catal. A Chem. 2009, 307, 77–87. [Google Scholar] [CrossRef]
  53. Long, W.; Brunelli, N.A.; Didas, S.A.; Ping, E.W.; Jones, C.W. Aminopolymer-Silica composite supported pd catalysts for selective hydrogenation of alkynes. ACS Catal. 2013, 3, 1700–1708. [Google Scholar] [CrossRef]
  54. Zhou, X.; Wu, T.; Hu, B.; Jiang, T.; Han, B. Ru nanoparticles stabilized by poly(N-vinyl-2-pyrrolidone) grafted onto silica: Very active and stable catalysts for hydrogenation of aromatics. J. Mol. Catal. A Chem. 2009, 306, 143–148. [Google Scholar] [CrossRef]
  55. Dzhardimalieva, G.I.; Uflyand, I.E. Preparation of metal-polymer nanocomposites by chemical reduction of metal ions: Functions of polymer matrices. J. Polym. Res. 2018, 25, 255. [Google Scholar] [CrossRef]
  56. Dzhardimalieva, G.I.; Zharmagambetova, A.K.; Kudaibergenov, S.E.; Uflyand, I.E. Polymer-immobilized clusters and metal nanoparticles in catalysis. Kinet. Catal. 2020, 61, 198–223. [Google Scholar] [CrossRef]
  57. Vögtle, F.; Richardt, G.; Werner, N. Dendrimer Chemistry: Concepts, Syntheses, Properties, Applications; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2009; p. 342. [Google Scholar]
  58. Newkome, G.R.; Moorefield, C.N.; Vögtle, F. Dendritic Macromolecules: Concepts, Synthesis, Perspectives; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 1996; p. 261. [Google Scholar] [CrossRef]
  59. Fréchet, J.M.J.; Tomalia, D. Dendrimers and Other Dendritic Polymers; John Wiley & Sons, Ltd.: Chichester, UK, 2001; p. 633. [Google Scholar] [CrossRef]
  60. Niu, Y.; Crooks, R.M. Dendrimer-encapsulated metal nanoparticles and their applications to catalysis. C. R. Chimie 2003, 6, 1049–1059. [Google Scholar] [CrossRef]
  61. Bosman, A.W.; Janssen, H.M.; Meijer, E.W. About Dendrimers: Structure, Physical Properties, and Applications. Chem. Rev. 1999, 99, 1665–1688. [Google Scholar] [CrossRef]
  62. Bronstein, L.M.; Shifrina, Z.B. Dendrimers as encapsulating, stabilizing, or directing agents for inorganic nanoparticles. Chem. Rev. 2011, 111, 5301–5344. [Google Scholar] [CrossRef]
  63. Olofsson, K.; Andren, O.C.J.; Malkochl, M. Recent advances on crosslinked dendritic networks. J. Appl. Polym. Sci. 2014, 131, 39876. [Google Scholar] [CrossRef]
  64. Tanabe, M.; Yamamoto, K. Precise synthesis of nanoparticles and their catalytic behavior. In Nanoparticles in Catalysis. Topics in Organometallic Chemistry; Kobayashi, S., Ed.; Springer: Cham, Switzerland, 2020; Volume 66, pp. 131–170. [Google Scholar] [CrossRef]
  65. Buhleier, E.; Wehner, W.; Vögtle, F. “Cascade” and “nonskid-chain-like” syntheses of molecular cavity topologies. Synthesis 1978, 1978, 155–158. [Google Scholar] [CrossRef]
  66. Fischer, M.; Vögtle, F. Dendrimers: From Design to Application—A Progress Report. Angew. Chem. Int. Ed. 1999, 38, 884–905. [Google Scholar] [CrossRef]
  67. Denkewalter, R.G.; Kolc, J.; Lukasavage, W.J. Macromolecular Highly Branched Homogeneous Compound Based on Lysine Units. U.S. Patent 4,289,872, 15 September 1981. [Google Scholar]
  68. Tomalia, D.A.; Baker, H.; Dewald, J.; Hall, M.; Kallos, G.; Martin, S.; Roeck, J.; Ryder, J.; Smith, P. A new class of polymers: Starburst-dendritic macromolecules. Polym. J. 1985, 17, 117–132. [Google Scholar] [CrossRef] [Green Version]
  69. Hawker, C.J.; Fréchet, J.M.J. Preparation of polymers with controlled molecular architecture. A new convergent approach to dendritic macromolecules. J. Am. Chem. Soc. 1990, 112, 7638–7647. [Google Scholar] [CrossRef]
  70. Newkome, G.R.; Yao, Z.-Q.; Baker, G.R.; Gupta, V.K. M.i.c.e.l.l.e.s. Part 1. Cascade molecules: A new approach to micelles. A [27]-arborol. J. Org. Chem. 1985, 50, 2003–2004. [Google Scholar] [CrossRef]
  71. Tomalia, D.A.; Nixon, L.S.; Hedstrand, D.M. The Role of branch cell symmetry and other critical nanoscale design parameters in the determination of dendrimer encapsulation properties. Biomolecules 2020, 10, 642. [Google Scholar] [CrossRef]
  72. Astruc, D.; Boisselier, E.; Ornelas, C. Dendrimers Designed for functions: From physical, photophysical, and supramolecular properties to applications in sensing, catalysis, molecular electronics, photonics, and nanomedicine. Chem. Rev. 2010, 110, 1857–1959. [Google Scholar] [CrossRef]
  73. Torres, C.C.; Campos, C.H.; Diáz, C.; Jiménez, V.A.; Vidal, F.; Guzmán, L.; Alderete, J.B. PAMAM-grafted TiO2 nanotubes as novel versatile materials for drug delivery applications. Mater. Sci. Eng. C 2016, 65, 164–171. [Google Scholar] [CrossRef]
  74. González, B.; Colilla, M.; López de Laorden, C.; Vallet-Regí, M. A novel synthetic strategy for covalently bonding dendrimers to ordered mesoporous silica: Potential drug delivery applications. J. Mater. Chem. 2009, 19, 9012–9024. [Google Scholar] [CrossRef]
  75. Svenson, S. The dendrimer paradox—High medical expectations but poor clinical translation. Chem. Soc. Rev. 2015, 44, 4131–4144. [Google Scholar] [CrossRef]
  76. Ramireddy, R.R.; Raghupathi, K.R.; Torres, D.A.; Thayumanavan, S. Stimuli sensitive amphiphilic dendrimers. New J. Chem. 2012, 36, 340–349. [Google Scholar] [CrossRef] [Green Version]
  77. Kono, K. Dendrimer-based bionanomaterials produced by surface modification, assembly and hybrid formation. Polym. J. 2012, 44, 531–540. [Google Scholar] [CrossRef] [Green Version]
  78. Wolinsky, J.; Grinstaff, M. Therapeutic and diagnostic applications of dendrimers for cancer treatment. Adv. Drug Deliv. Rev. 2008, 60, 1037–1055. [Google Scholar] [CrossRef]
  79. Astruc, D. Electron-transfer processes in dendrimers and their implication in biology, catalysis, sensing and nanotechnology. Nat. Chem. 2012, 4, 255–267. [Google Scholar] [CrossRef]
  80. Yamamoto, K.; Imaoka, T. Dendrimer complexes based on fine-controlled metal assembling. Bull. Chem. Soc. Jpn. 2006, 79, 511–526. [Google Scholar] [CrossRef]
  81. Vögtle, F.; Gestermann, S.; Hesse, R.; Schwierz, H.; Windisch, B. Functional dendrimers. Prog. Polym. Sci. 2000, 25, 987–1041. [Google Scholar] [CrossRef]
  82. Newkome, G.R.; Shreiner, C.D. Poly(amidoamine), polypropylenimine, and related dendrimers and dendrons possessing different 1 → 2 branching motifs: An overview of the divergent procedures. Polymer 2008, 49, 1–173. [Google Scholar] [CrossRef] [Green Version]
  83. Newkome, G.R.; Shreiner, C. Dendrimers derived from 1 → 3 branching motifs. Chem. Rev. 2010, 110, 6338–6442. [Google Scholar] [CrossRef]
  84. Smith, D.K.; Hirst, A.R.; Love, C.S.; Hardy, J.G.; Brignell, S.V.; Huang, B. Self-assembly using dendritic building blocks—Towards controllable nanomaterials. Prog. Polym. Sci. 2005, 30, 220–293. [Google Scholar] [CrossRef] [Green Version]
  85. Tomalia, D.A. Dendrons/dendrimers: Quantized, nano-element like building blocks for soft-soft and soft-hard nano-compound synthesis. Soft Matter. 2010, 6, 456–474. [Google Scholar] [CrossRef]
  86. Peleshanko, S.; Tsukruk, V.V. The architectures and surface behavior of highly branched molecules. Prog. Polym. Sci. 2008, 33, 523–580. [Google Scholar] [CrossRef]
  87. Caminade, A.-M.; Ouali, A.; Keller, M.; Majoral, J.-P. Organocatalysis with dendrimers. Chem. Soc. Rev. 2012, 41, 4113–4125. [Google Scholar] [CrossRef]
  88. Peng, X.; Pan, Q.; Rempel, G.L. Bimetallic dendrimer-encapsulated nanoparticles as catalysts: A review of the research advances. Chem. Soc. Rev. 2008, 37, 1619–1628. [Google Scholar] [CrossRef]
  89. van Heerbeek, R.; Kamer, P.C.J.; van Leeuwen, P.W.N.M.; Reek, J.N.H. Dendrimers as support for recoverable catalysts and reagents. Chem. Rev. 2002, 102, 3717–3756. [Google Scholar] [CrossRef]
  90. Wang, D.; Astruc, D. Dendritic catalysis—Basic concepts and recent trends. Coord. Chem. Rev. 2013, 257, 2317–2334. [Google Scholar] [CrossRef]
  91. Scott, R.W.J.; Wilson, O.M.; Crooks, R.M. Synthesis, characterization, and applications of dendrimer-encapsulated nanoparticles. J. Phys. Chem. B 2005, 109, 692–704. [Google Scholar] [CrossRef]
  92. Chung, Y.-M.; Rhee, H.-K. Internal/external use of dendrimer in catalysis. Korean J. Chem. Eng. 2004, 21, 81–97. [Google Scholar] [CrossRef]
  93. Andrés, R.; de Jesús, E.; Flores, J.C. Catalysts based on palladium dendrimers. New J. Chem. 2007, 31, 1161–1191. [Google Scholar] [CrossRef]
  94. Van de Coevering, R.; Klein Gebbink, R.J.M.; van Koten, G. Soluble organic supports for the non-covalent immobilization of homogeneous catalysts; modular approaches towards sustainable catalysts. Prog. Polym. Sci. 2005, 30, 474–490. [Google Scholar] [CrossRef]
  95. Chandler, B.D.; Gilbertson, J.D. Dendrimer-Encapsulated bimetallic nanoparticles: Synthesis, characterization, and applications to homogeneous and heterogeneous catalysis. In Dendrimer Catalysis. Topics in Organometallic Chemistry; Gade, L.H., Ed.; Springer: Berlin/Heidelberg, Germany, 2006; Volume 20, pp. 97–120. [Google Scholar] [CrossRef]
  96. Tomalia, D.A.; Dewald, J.R. Dense Star Polymers Having Core, Core Branches, Terminal Groups. U.S. Patent 4,507,466 A, 26 March 1985. [Google Scholar]
  97. De Brabander-van den Berg, E.M.M.; Meijer, E.W. Poly(propylene imine) Dendrimers: Large-scale synthesis by hetereogeneously catalyzed hydrogenations. Angew. Chem. Int. Ed. 1993, 32, 1308–1312. [Google Scholar] [CrossRef] [Green Version]
  98. Krishna, T.R.; Jain, S.; Tatu, U.S.; Jayaraman, N. Synthesis and biological evaluation of 3-amino-propan-1-ol based poly(ether imine) dendrimers. Tetrahedron 2005, 61, 4281–4288. [Google Scholar] [CrossRef]
  99. Jayamurugan, G.; Jayaraman, N. Synthesis of large generation poly(propyl ether imine) (PETIM) dendrimers. Tetrahedron 2006, 62, 9582–9588. [Google Scholar] [CrossRef]
  100. Krishna, T.R.; Jayaraman, N. Synthesis of poly(propylether imine) dendrimers and evaluation of their cytotoxic properties. J. Org. Chem. 2003, 68, 9694–9704. [Google Scholar] [CrossRef]
  101. Krishna, T.R.; Jayaraman, N. Synthesis and catalytic activities of Pd (II) phosphine complexes modified poly(ether imine) dendrimers. Tetrahedron 2004, 60, 10325–10334. [Google Scholar] [CrossRef]
  102. Jayamurugan, G.; Umesh, C.P.; Jayaraman, N. Preparation and catalytic studies of palladium nanoparticles stabilized by dendritic phosphine ligand-functionalized silica. J. Mol. Catal. A Chem. 2009, 307, 142–148. [Google Scholar] [CrossRef]
  103. Newkome, G.R.; Behera, R.K.; Moorefield, C.N.; Baker, G.R. Chemistry of micelles. 18. Cascade polymers: Syntheses and characterization of one-directional arborols based on adamantane. J. Org. Chem. 1991, 56, 7162–7167. [Google Scholar] [CrossRef]
  104. Hawker, C.; Fréchet, J.M.J. A new convergent approach to monodisperse dendritic macromolecules. Chem. Commun. 1990, 1010–1013. [Google Scholar] [CrossRef]
  105. Ihre, H.; Padilla De Jesus, O.L.; Fréchet, J.M.J. Fast and Convenient divergent synthesis of aliphatic ester dendrimers by anhydride coupling. J. Am. Chem. Soc. 2001, 123, 5908–5917. [Google Scholar] [CrossRef]
  106. Krska, S.W.; Seyferth, D. Synthesis of water-soluble carbosilane dendrimers. J. Am. Chem. Soc. 1998, 120, 3604–3612. [Google Scholar] [CrossRef]
  107. Ruiz, J.; Lafuente, G.; Marcen, S.; Ornelas, C.; Lazare, S.; Cloutet, E.; Blais, J.-C.; Astruc, D. Construction of giant dendrimers using a tripodal building block. J. Am. Chem. Soc. 2003, 125, 7250–7257. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, D.; Denux, D.; Ruiz, J.; Astruc, D. The clicked pyridyl-triazole ligand: From homogeneous to robust, recyclable heterogeneous mono- and polymetallic palladium catalysts for efficient suzuki–miyaura, sonogashira, and heck reactions. Adv. Synth. Catal. 2013, 355, 129–142. [Google Scholar] [CrossRef]
  109. Imaoka, T.; Kawana, Y.; Yamamoto, K. Copper-containing bimetallic complexes as efficient catalysts for aerobic oxidative coupling polymerization. Polym. Adv. Technol. 2011, 22, 1261–1265. [Google Scholar] [CrossRef]
  110. Caminade, A.M.; Majoral, J.P. Nanomaterials based on phosphorus dendrimers. Acc. Chem. Res. 2004, 37, 341–348. [Google Scholar] [CrossRef]
  111. Caminade, A.M.; Majoral, J.P. Phosphorus dendrimers possessing metallic groups in their internal structure (core or branches): Syntheses and properties. Coord. Chem. Rev. 2005, 249, 1917–1926. [Google Scholar] [CrossRef]
  112. Caminade, A.M.; Maraval, A.; Majoral, J.P. Phosphorus-containing dendrons: Synthesis, reactivity, properties, and use as building blocks for various dendritic architectures. Eur. J. Inorg. Chem. 2006, 2006, 887–901. [Google Scholar] [CrossRef]
  113. Caminade, A.-M.; Laurent, R. Homogeneous catalysis with phosphorus dendrimer complexes. Coord. Chem. Rev. 2019, 389, 59–72. [Google Scholar] [CrossRef]
  114. Tomalia, D.A.; Fréchet, J.M.J. Discovery of dendrimers and dendritic polymers: A brief historical perspective. J. Polym. Sci. A Polym. Chem. 2002, 40, 2719–2728. [Google Scholar] [CrossRef]
  115. Crooks, R.M.; Zhao, M.; Sun, L.; Chechik, V.; Yeung, L.K. Dendrimer-encapsulated metal nanoparticles: Synthesis, characterization, and applications to catalysis. Acc. Chem. Res. 2001, 34, 181–190. [Google Scholar] [CrossRef] [Green Version]
  116. Ooe, M.; Murata, M.; Mizugaki, T.; Ebitani, T.; Kaneda, K. Dendritic nanoreactors encapsulating Pd particles for substrate-specific hydrogenation of olefins. Nano Lett. 2002, 2, 999–1002. [Google Scholar] [CrossRef]
  117. Oh, S.-K.; Niu, Y.; Crooks, R.M. Size-selective catalytic activity of Pd nanoparticles encapsulated within end-group functionalized dendrimers. Langmuir 2005, 21, 10209–10213. [Google Scholar] [CrossRef]
  118. Pittelkow, M.; Moth-Poulsen, K.; Boas, U.; Christensen, J.B. Poly(amidoamine)-dendrimer-stabilized Pd (0) Nanoparticles as a catalyst for the Suzuki reaction. Langmuir 2003, 19, 7682–7684. [Google Scholar] [CrossRef]
  119. Deng, G.-J.; Fan, Q.-H.; Chen, X.-M.; Liu, D.-S.; Chan, A.S.C. A novel system consisting of easily recyclable dendritic Ru-BINAP catalyst for asymmetric hydrogenation. Chem. Commun. 2002, 15, 1570–1571. [Google Scholar] [CrossRef]
  120. Fan, Q.-H.; Chen, Y.-M.; Chen, X.-M.; Jiang, D.-Z.; Xi, F.; Chan, A.S.C. Highly effective and recyclable dendritic BINAP ligands for asymmetric hydrogenation. Chem. Commun. 2000, 9, 789–790. [Google Scholar] [CrossRef]
  121. Gopidas, K.R.; Whitesell, J.K.; Fox, M.A. Synthesis, characterization, and catalytic applications of a palladium-nanoparticle-cored dendrimer. Nano Lett. 2003, 3, 1757–1760. [Google Scholar] [CrossRef]
  122. Maximov, A.; Zolotukhina, A.; Murzin, V.; Karakhanov, E.; Rosenberg, E. Ruthenium nanoparticles stabilized in cross-linked dendrimer matrices: Hydrogenation of phenols in aqueous media. ChemCatChem 2015, 7, 1197–1210. [Google Scholar] [CrossRef]
  123. Karakhanov, E.A.; Maksimov, A.L.; Zakharian, E.M.; Kardasheva, Y.S.; Savilov, S.V.; Truhmanova, N.I.; Ivanov, A.O.; Vinokurov, V.A. Palladium nanoparticles encapsulated in a dendrimer networks as catalysts for the hydrogenation of unsaturated hydrocarbons. J. Mol. Catal. A Chem. 2015, 397, 1–18. [Google Scholar] [CrossRef]
  124. Karakhanov, E.A.; Maximov, A.L.; Tarasevich, B.N.; Skorkin, V.A. Dendrimer-based catalysts in Wacker-oxidation: Unexpected selectivity to terminal double bonds. J. Mol. Catal. A Chem. 2009, 297, 73–79. [Google Scholar] [CrossRef]
  125. Bernechea, M.; de Jesús, E.; López-Mardomingo, C.; Terreros, P. Dendrimer-encapsulated Pd nanoparticles versus palladium acetate as catalytic precursors in the stille reaction in water. Inorg. Chem. 2009, 48, 4491–4496. [Google Scholar] [CrossRef]
  126. Borkowski, T.; Subik, P.; Trzeciak, A.M.; Wołowiec, S. Palladium (0) deposited on PAMAM dendrimers as a catalyst for C–C cross coupling reactions. Molecules 2011, 16, 427–441. [Google Scholar] [CrossRef] [Green Version]
  127. Karakhanov, E.; Maximov, A.; Zolotukhina, A.; Kardasheva, Y.; Talanova, M. Thermo-responsive ruthenium dendrimer-based catalysts for hydrogenation of the aromatic compounds and phenols. J. Inorg. Organomet. Polym. Mater. 2016, 26, 1264–1279. [Google Scholar] [CrossRef]
  128. Karakhanov, E.A.; Maximov, A.L.; Zakharyan, E.M.; Zolotukhina, A.V.; Ivanov, A.O. Palladium nanoparticles on dendrimer-containing supports as catalysts for hydrogenation of unsaturated hydrocarbons. Mol. Catal. 2017, 440, 107–119. [Google Scholar] [CrossRef]
  129. Wu, L.; Li, B.-L.; Huang, Y.-Y.; Zhou, H.-F.; He, Y.-M.; Fan, Q.-H. Phosphine dendrimer-stabilized palladium nanoparticles, a highly active and recyclable catalyst for the suzuki−miyaura reaction and hydrogenation. Org. Lett. 2006, 8, 3605–3608. [Google Scholar] [CrossRef]
  130. Keller, M.; Hameau, A.; Spataro, G.; Ladeira, S.; Caminade, A.-M.; Majoral, J.-P.; Ouali, A. An efficient and recyclable dendritic catalyst able to dramatically decrease palladium leaching in Suzuki coupling. Green Chem. 2012, 14, 2807–2815. [Google Scholar] [CrossRef]
  131. Murata, M.; Tanaka, Y.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Palladium–platinum bimetallic nanoparticle catalysts using dendron assembly for selective hydrogenation of dienes and their application to thermomorphic system. Chem. Lett. 2005, 34, 272–273. [Google Scholar] [CrossRef]
  132. Ooe, M.; Murata, M.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Supramolecular catalysts by encapsulating palladium complexes within dendrimers. J. Am. Chem. Soc. 2004, 126, 1604–1605. [Google Scholar] [CrossRef]
  133. Kimura, M.; Kato, M.; Muto, T.; Hanabusa, K.; Shirai, H. Temperature-sensitive dendritic hosts: Synthesis, characterization, and control of catalytic activity. Macromolecules 2000, 33, 1117–1119. [Google Scholar] [CrossRef]
  134. Ramírez-Crescencio, F.; Enciso, A.E.; Hasan, M.; Da Costa, V.C.P.; Annunziata, O.; Redón, R.; Coffer, J.L.; Simanek, E.E. Thermoregulated coacervation, metal-encapsulation and nanoparticle synthesis in novel triazine dendrimers. Molecules 2016, 21, 599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Zheng, K.; Ren, J.; He, J. Thermally responsive unimolecular nanoreactors from amphiphilic dendrimer-like copolymer prepared via anionic polymerization and cross metathesis reaction. Macromolecules 2019, 52, 6780–6791. [Google Scholar] [CrossRef]
  136. Mery, D.; Astruc, D. Dendritic catalysis: Major concepts and recent progress. Coord. Chem. Rev. 2006, 250, 1965–1979. [Google Scholar] [CrossRef]
  137. Ratheesh, V.K.; Karical, K.; Gopidas, R. Palladium nanoparticle-cored G1-dendrimer stabilized by carbon–Pd bonds: Synthesis, characterization and use as chemoselective, room temperature hydrogenation catalyst. Tetrahedron Lett. 2011, 52, 3102–3105. [Google Scholar] [CrossRef]
  138. Nlate, S.; Astruc, D.; Neumann, R. Synthesis, catalytic activity in oxidation reactions, and recyclability of stable polyoxometalate-centred dendrimers. Adv. Synth. Catal. 2004, 346, 1445–1448. [Google Scholar] [CrossRef]
  139. Kakkar, A.K. Dendritic polymers: From efficient catalysis to drug delivery. Macromol. Symp. 2003, 196, 145–154. [Google Scholar] [CrossRef]
  140. Mizugaki, T.; Ooe, M.; Ebitani, K.; Kaneda, K. Catalysis of dendrimer-bound Pd (II) complex selective hydrogenation of conjugated dienes to monoenes. J. Mol. Catal. A Chem. 1999, 145, 329–333. [Google Scholar] [CrossRef]
  141. Reetz, M.T.; Lohmer, G.; Schwickardi, R. Synthesis and catalytic activity of dendritic diphosphane metal complexes. Angew. Chem. Int. Ed. 1997, 36, 1526–1529. [Google Scholar] [CrossRef]
  142. Heuzé, K.; Méry, D.; Gauss, D.; Astruc, D. Copper-free, recoverable dendritic Pd catalysts for the Sonogashira reaction. Chem. Commun. 2003, 2274–2275. [Google Scholar] [CrossRef]
  143. Smith, G.S.; Mapolie, S.F. Iminopyridyl-palladium dendritic catalyst precursors: Evaluation in Heck reactions. J. Mol. Catal. A Chem. 2004, 213, 187–192. [Google Scholar] [CrossRef]
  144. Catsoulacos, D.P.; Steele, B.R.; Heropoulos, G.A.; Micha-Screttas, M.; Screttas, C.G. An iminophosphine dendrimeric ligand and its evaluation in the Heck reaction. Tetrahedron Lett. 2003, 44, 4575–4578. [Google Scholar] [CrossRef]
  145. Astruc, D.; Heuzé, K.; Gatard, S.; Méry, D.; Nlate, S.; Plault, L. Metallodendritic catalysis for redox and carbon–carbon bond formation reactions: A step towards green chemistry. Adv. Synth. Catal. 2005, 347, 329–338. [Google Scholar] [CrossRef]
  146. Bourque, S.C.; Maltais, F.; Xiao, W.-J.; Tardif, O.; Alper, H.; Arya, P.; Manzer, L.E. Hydroformylation reactions with rhodium-complexed dendrimers on silica. J. Am. Chem. Soc. 1999, 121, 3035–3038. [Google Scholar] [CrossRef] [Green Version]
  147. Zweni, P.P.; Alper, H. Dendrimer-palladium complex catalyzed oxidation of terminal alkenes to methyl ketones. Adv. Synth. Catal. 2004, 346, 849–854. [Google Scholar] [CrossRef]
  148. Zweni, P.P.; Alper, H. Silica-supported dendrimer-palladium complex-catalyzed selective hydrogenation of dienes to monoolefins. Adv. Synth. Catal. 2006, 348, 725–731. [Google Scholar] [CrossRef]
  149. Reynhardt, J.P.K.; Alper, H. Hydroesterification reactions with palladium-complexed PAMAM dendrimers immobilized on silica. J. Org. Chem. 2003, 68, 22, 8353–8360. [Google Scholar] [CrossRef] [PubMed]
  150. Ornelas, C.; Ruiz Aranzaes, J.; Salmon, L.; Astruc, D. “Click” dendrimers: Synthesis, redox sensing of Pd(OAc)2, and remarkable catalytic hydrogenation activity of precise pd nanoparticles stabilized by 1,2,3-triazole-containing dendrimers. Chem. Eur. J. 2008, 14, 50–64. [Google Scholar] [CrossRef]
  151. Niu, Y.; Crooks, R.M. Preparation of dendrimer-encapsulated metal nanoparticles using organic solvents. Chem. Mater. 2003, 15, 3463–3467. [Google Scholar] [CrossRef]
  152. Zhao, M.; Crooks, R.M. Homogeneous hydrogenation catalysis with monodisperse, dendrimer-encapsulated Pd and Pt Nanoparticles. Angew. Chem. Int. Ed. 1999, 38, 364–366. [Google Scholar] [CrossRef]
  153. Zhao, M.; Crooks, R.M. Dendrimer-encapsulated Pt nanoparticles: Synthesis, characterization, and applications to catalysis. Adv. Mater. 1999, 11, 217–220. [Google Scholar] [CrossRef]
  154. Myers, V.S.; Weier, M.G.; Carino, E.V.; Yancey, D.F.; Pande, S.; Crooks, R.M. Dendrimer-encapsulated nanoparticles: New synthetic and characterization methods and catalytic applications. Chem. Sci. 2011, 2, 1632–1646. [Google Scholar] [CrossRef] [Green Version]
  155. Mizugaki, T.; Kibata, T.; Ota, K.; Mitsudome, T.; Ebitani, K.; Jitsukawa, K.; Kaneda, K. Controlled synthesis of Pd clusters in subnanometer range using poly(propylene imine) dendrimers. Chem. Lett. 2009, 38, 1118–1119. [Google Scholar] [CrossRef]
  156. Maeno, Z.; Kibata, T.; Mitsudome, T.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Subnanoscale size effect of dendrimer-encapsulated pd clusters on catalytic hydrogenation of olefin. Chem. Lett. 2011, 40, 180–181. [Google Scholar] [CrossRef]
  157. Kibata, T.; Mitsudome, T.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Investigation of size-dependent properties of sub-nanometer palladium clusters encapsulated within a polyamine dendrimer. Chem. Commun. 2013, 49, 167–169. [Google Scholar] [CrossRef]
  158. Zhang, L.; Anderson, R.M.; Crooks, R.M.; Henkelman, G. Correlating structure and function of metal nanoparticles for catalysis. Surf. Sci. 2015, 640, 65–72. [Google Scholar] [CrossRef] [Green Version]
  159. Chung, Y.-M.; Rhee, H.-K. Partial hydrogenation of 1,3-cyclooctadiene using dendrimer-encapsulated Pd–Rh bimetallic nanoparticles. J. Mol. Catal. A Chem. 2003, 206, 291–298. [Google Scholar] [CrossRef]
  160. Chung, Y.-M.; Rhee, H.-K. Pt–Pd bimetallic nanoparticles encapsulated in dendrimer nanoreactor. Catal. Lett. 2003, 85, 159–164. [Google Scholar] [CrossRef]
  161. Scott, R.W.J.; Datye, A.K.; Crooks, R.M. Bimetallic palladium−platinum dendrimer-encapsulated catalysts. J. Am. Chem. Soc. 2003, 125, 3708–3709. [Google Scholar] [CrossRef]
  162. Scott, R.W.J.; Wilson, O.M.; Oh, S.-K.; Kenik, E.A.; Crooks, R.M. Bimetallic Palladium−gold dendrimer-encapsulated catalysts. J. Am. Chem. Soc. 2004, 126, 15583–15591. [Google Scholar] [CrossRef]
  163. Peng, X.; Pan, Q.; Rempel, G.L.; Wu, S. Synthesis, characterization, and application of PdPt and PdRh bimetallic nanoparticles encapsulated within amine-terminated poly(amidoamine) dendrimers. Catal. Commun. 2009, 11, 62–66. [Google Scholar] [CrossRef]
  164. Wang, Y.; Peng, X. RuRh bimetallic nanoparticles stabilized by 15-membered macrocycles-terminated poly(propylene imine) dendrimer: Preparation and catalytic hydrogenation of nitrile–butadiene rubber. Nano-Micro Lett. 2014, 6, 55–62. [Google Scholar] [CrossRef]
  165. Luo, L.; Duan, Z.; Li, H.; Kim, J.; Henkelman, G.; Crooks, R.M. Tunability of the Adsorbate binding on bimetallic alloy nanoparticles for the optimization of catalytic hydrogenation. J. Am. Chem. Soc. 2017, 139, 5538–5546. [Google Scholar] [CrossRef]
  166. Nakamula, I.; Yamanoi, Y.; Imaoka, T.; Yamamoto, K.; Nishihara, H. A Uniform bimetallic rhodium/iron nanoparticle catalyst for the hydrogenation of olefins and nitroarenes. Angew. Chem. Int. Ed. 2011, 50, 5830–5833. [Google Scholar] [CrossRef]
  167. Wilson, O.M.; Scott, R.W.J.; Garcia-Martinez, J.C.; Crooks, R.M. Synthesis, Characterization, and structure-selective extraction of 1−3-nm diameter AuAg dendrimer-encapsulated bimetallic nanoparticles. J. Am. Chem. Soc. 2005, 127, 1015–1024. [Google Scholar] [CrossRef]
  168. Endo, T.; Yoshimura, T.; Esumi, K. Synthesis and catalytic activity of gold–silver binary nanoparticles stabilized by PAMAM dendrimer. J. Colloid Interf. Sci. 2005, 286, 602–609. [Google Scholar] [CrossRef]
  169. Lang, H.; Maldonado, S.; Stevenson, K.J.; Chandler, B.D. Synthesis and characterization of dendrimer templated supported bimetallic Pt−Au nanoparticles. J. Am. Chem. Soc. 2004, 126, 12949–12956. [Google Scholar] [CrossRef] [Green Version]
  170. Marvin, K.A.; Thadani, N.N.; Atkinson, C.A.; Keller, E.L.; Stevenson, K.J. Preparation and catalytic evaluation of ruthenium–nickel dendrimer encapsulated nanoparticles via intradendrimer redox displacement of nickel nanoparticles. Chem. Commun. 2012, 48, 6289–6291. [Google Scholar] [CrossRef] [Green Version]
  171. Karakhanov, E.; Maximov, A.; Kardasheva, Y.; Semernina, V.; Zolotukhina, A.; Ivanov, A.; Abbott, G.; Rosenberg, E.; Vinokurov, V. Pd nanoparticles in dendrimers immobilized on silica−polyamine composites as catalysts for selective hydrogenation. ACS Appl. Mater. Interfaces 2014, 6, 8807–8816. [Google Scholar] [CrossRef]
  172. Kliger, E.G.; Shuikina, L.P.; Parenago, O.P.; Frolov, V.M. Homogeneous and supported platinum complex catalysts with nitrogen-containing ligands in the hydrogenation of unsaturated hydrocarbons. Russ. Chem Bull. 1989, 37, 2182–2184. [Google Scholar] [CrossRef]
  173. Turisbekova, K.L.; Shuikina, L.P.; Parenago, O.P.; Frolov, V.M. Complexes of ruthenium and rhodium with aliphatic amines in the catalysis of hydrogenation of unsaturated hydrocarbons. Kinet. Catal. 1988, 29, 885. [Google Scholar]
  174. Frolov, V.M. Platinum metals complex catalysts for liquid-phase hydrogenations novel catalysts utilising aliphatic amines and related systems. Platinum Metals Rev. 1996, 40, 8–18. [Google Scholar]
  175. Frolov, V.M.; Parenago, O.P.; Bondarenko, G.N.; Kovaleva, L.S.; Elnatanova, A.I.; Shuikina, L.P.; Cherkashin, G.M.; Mirskaya, E.Y. Catalysts for the selective hydrogenation of conjugated dienes and acetylenes to olefins obtained by reducing palladium salts BY tertiary-amines. Kinet. Katal. 1981, 22, 1356. [Google Scholar]
  176. Frolov, V.M.; Parenago, O.P.F. New palladium-complex catalysts for selective hydrogenation. D. I. Mendeleev All-Union Chem. Soc. J. 1989, 34, 659–665. (In Russian) [Google Scholar]
  177. Macdonald, R.P.; Winterbottom, J.M. Palladium-polyamide catalysts for the hydrogenation of some unsaturated hydrocarbons in the liquid phase. J. Catal. 1979, 57, 195–207. [Google Scholar] [CrossRef]
  178. Takahashi, M.; Imaoka, T.; Hongo, Y.; Yamamoto, K. Formation of a Pt12 Cluster by single-atom control that leads to enhanced reactivity: Hydrogenation of unreactive olefins. Angew. Chem. Int. Ed. 2013, 52, 7419–7421. [Google Scholar] [CrossRef] [PubMed]
  179. Takahashi, M.; Imaoka, T.; Hongo, Y.; Yamamoto, K. A highly-active and poison-tolerant Pt12 sub-nanocluster catalyst for the reductive amination of aldehydes with amines. Dalton Trans. 2013, 42, 15919–15921. [Google Scholar] [CrossRef] [PubMed]
  180. Huda, M.; Minamisawa, K.; Tsukamoto, T.; Tanabe, M.; Yamamoto, K. Aerobic toluene oxidation catalyzed by subnano metal particles. Angew. Chem. Int. Ed. 2019, 58, 1002–1006. [Google Scholar] [CrossRef]
  181. Kambe, T.; Haruta, N.; Imaoka, T.; Yamamoto, K. Solution phase synthesis of Al13 using a dendrimer template. Nat. Commun. 2017, 8, 2046. [Google Scholar] [CrossRef] [Green Version]
  182. Kambe, T.; Imaoka, T.; Yamamoto, K. Reducing capsule based on electron programming: Versatile synthesizer for size-controlled ultra-small metal clusters. Chem. Eur. J. 2016, 22, 16406–16409. [Google Scholar] [CrossRef]
  183. Takahashi, M.; Koizumi, H.; Chun, W.-J.; Kori, M.; Imaoka, T.; Yamamoto, K. Finely controlled multimetallic nanocluster catalysts for solvent-free aerobic oxidation of hydrocarbons. Sci. Adv. 2017, 3, e1700101–e1700109. [Google Scholar] [CrossRef] [Green Version]
  184. Tsukamoto, T.; Kambe, T.; Nakao, A.; Imaoka, T.; Yamamoto, K. Atom-hybridization for synthesis of polymetallic clusters. Nat. Commun. 2018, 9, 3873. [Google Scholar] [CrossRef] [Green Version]
  185. Takanashi, K.; Fujii, A.; Nakajima, R.; Chiba, H.; Higuchi, M.; Einaga, Y.; Yamamoto, K. Heterometal assembly in dendritic polyphenylazomethines. Bull. Chem. Soc. Jpn. 2007, 80, 1563–1572. [Google Scholar] [CrossRef]
  186. Deraedt, C.; Salmon, L.; Astruc, D. “Click” dendrimer-stabilized palladium nanoparticles as a green catalyst down to parts per million for efficient c–c cross-coupling reactions and reduction of 4-nitrophenol. Adv. Synth. Catal. 2014, 356, 2525–2538. [Google Scholar] [CrossRef]
  187. Li, N.; Echeverría, M.; Moya, S.; Ruiz, J.; Astruc, D. “Click” synthesis of nona-PEG-branched triazole dendrimers and stabilization of gold nanoparticles that efficiently catalyze p-nitrophenol reduction. Inorg. Chem. 2014, 53, 6954–6961. [Google Scholar] [CrossRef]
  188. Deraedt, C.; Pinaud, N.; Astruc, D. Recyclable catalytic dendrimer nanoreactor for part-per-million CuI catalysis of “click” chemistry in water. J. Am. Chem. Soc. 2014, 136, 12092–12098. [Google Scholar] [CrossRef]
  189. Ornelas, C.; Ruiz, J.; Salmon, L.; Astruc, D. Sulphonated “click” dendrimer-stabilized palladium nanoparticles as highly efficient catalysts for olefin hydrogenation and suzuki coupling reactions under ambient conditions in aqueous media. Adv. Synth. Catal. 2008, 350, 837–845. [Google Scholar] [CrossRef]
  190. Liang, L.; Ruiz, J.; Astruc, D. The efficient copper (I) (Hexabenzyl)tren Catalyst and dendritic analogues for green “click” reactions between azides and alkynes in organic solvent and in water: Positive dendritic effects and monometallic mechanism. Adv. Synth. Catal. 2011, 353, 3434–3450. [Google Scholar] [CrossRef]
  191. Liu, X.; Gregurec, D.; Irigoyen, J.; Martinez, A.; Moya, S.; Ciganda, R.; Hermange, P.; Ruiz, J.; Astruc, D. Precise localization of metal nanoparticles in dendrimer nanosnakes or inner periphery and consequences in catalysis. Nat. Commun. 2016, 7, 13152–13159. [Google Scholar] [CrossRef]
  192. Esumi, K.; Isono, R.; Yoshimura, T. Preparation of PAMAM and PPI-metal (silver, platinum, and palladium) nanocomposites and their catalytic activities for reduction of 4-nitrophenol. Langmuir 2004, 20, 237–243. [Google Scholar] [CrossRef]
  193. Karakanov, E.A.; Zolotukhina, A.V.; Ivanov, A.O.; Maximov, A.L. Dendrimer-encapsulated pd nanoparticles, immobilized in silica pores, as catalysts for selective hydrogenation of unsaturated compounds. ChemistryOpen 2019, 8, 358–381. [Google Scholar] [CrossRef]
  194. Cutler, E.C.; Lundin, E.; Garabato, B.D.; Choi, D.; Shon, Y.-S. Dendritic functionalization of monolayer-protected gold nanoparticles. Mater. Res. Bull. 2007, 42, 1178–1185. [Google Scholar] [CrossRef]
  195. Gopidas, K.R.; Whitesell, J.K.; Fox, M.A. Metal-core−organic shell dendrimers as unimolecular micelles. J. Am. Chem. Soc. 2003, 125, 14168–14180. [Google Scholar] [CrossRef]
  196. Gopidas, K.R.; Whitesell, J.K.; Fox, M.A. Nanoparticle-cored dendrimers:  synthesis and characterization. J. Am. Chem. Soc. 2003, 125, 6491–6502. [Google Scholar] [CrossRef]
  197. Wang, R.Y.; Yang, J.; Zheng, Z.P.; Carducci, M.D.; Jiao, J.; Seraphin, S. Dendron-controlled nucleation and growth of gold nanoparticles. Angew. Chem. Int. Ed. 2001, 40, 549–552. [Google Scholar] [CrossRef]
  198. Kim, M.K.; Jeon, Y.M.; Jeon, W.S.; Kim, H.J.; Hong, S.G.; Park, C.G.; Kim, K. Novel dendron-stabilized gold nanoparticles with high stability and narrow size distribution. Chem. Commun. 2001, 7, 667–668. [Google Scholar] [CrossRef] [Green Version]
  199. Jiang, G.H.; Wang, L.; Chen, T.; Yu, H.J.; Chen, C. Preparation of gold nanoparticles in the presence of poly(benzyl ether) alcohol dendrons. Mater. Chem. Phys. 2006, 98, 76–82. [Google Scholar] [CrossRef]
  200. Yang, P.; Zhang, W.; Du, Y.; Wang, X. Hydrogenation of nitrobenzenes catalyzed by platinum nanoparticle core-polyaryl ether trisacetic acid ammonium chloride dendrimer shell nanocomposite. J. Mol. Catal. A Chem. 2006, 260, 4–10. [Google Scholar] [CrossRef]
  201. Shon, Y.-S.; Choi, D.; Dare, J.; Dinh, T. Synthesis of nanoparticle-cored dendrimers by convergent dendritic functionalization of monolayer-protected nanoparticles. Langmuir 2008, 24, 6924–6931. [Google Scholar] [CrossRef]
  202. Sutton, A.; Franc, G.; Kakkar, A. Silver metal nanoparticles: Facile dendrimer-assisted size-controlled synthesis and selective catalytic reduction of chloronitrobenzenes. J. Polym. Sci. A Polym. Chem. 2009, 47, 4482–4493. [Google Scholar] [CrossRef]
  203. Zhang, W.; Li, L.; Du, Y.; Wang, X.; Yang, P. Gold/platinum bimetallic core/shell nanoparticles stabilized by a fréchet-type dendrimer: Preparation and catalytic hydrogenations of phenylaldehydes and nitrobenzenes. Catal. Lett. 2009, 127, 429–436. [Google Scholar] [CrossRef]
  204. Love, C.S.; Chechik, V.; Smith, D.K.; Brennan, C. Dendron-stabilised gold nanoparticles: Generation dependence of core size and thermal stability. J. Mater. Chem. 2004, 14, 919–923. [Google Scholar] [CrossRef]
  205. Wang, Y.A.; Li, J.J.; Chen, H.Y.; Peng, X.G. Stabilization of inorganic nanocrystals by organic dendrons. J. Am. Chem. Soc. 2002, 124, 2293–2298. [Google Scholar] [CrossRef]
  206. Guo, W.H.; Li, J.J.; Wang, Y.A.; Peng, X.G. Luminescent CdSe/CdS core/shell nanocrystals in dendron boxes: Superior chemical, photochemical and thermal stability. J. Am. Chem. Soc. 2003, 125, 3901–3909. [Google Scholar] [CrossRef]
  207. Mizugaki, T.; Muratra, M.; Fukubayashi, S.; Mitsudome, T.; Jitsujkawa, K.; Kaneda, K. PAMAM dendron-stabilised palladium nanoparticles: Effect of generation and peripheral groups on particle size and hydrogenation activity. Chem. Commun. 2008, 2, 241–243. [Google Scholar] [CrossRef]
  208. Gryzdova, D.V.; Smirnova, N.S.; Shlyapin, D.A.; Tsyrul’nikov, P.G. Gas-phase and liquid-phase hydrogenation of acetylene in lean and enriched mixtures over supported modified palladium catalysts. Russ. Chem. J. 2018, 62, 89–109. (In Russian) [Google Scholar]
  209. Mashkovsky, I.S.; Baeva, A.G.N.; Stakheev, Y.; Vargaftik, M.N.; Kozitsyna, N.Y.; Moiseev, I.I. Novel Pd–Zn/C catalyst for selective alkyne hydrogenation: Evidence for the formation of Pd–Zn bimetallic alloy particles. Mendeleev Commun. 2014, 24, 355–357. [Google Scholar] [CrossRef]
  210. Stakheev, A.Y.; Moroz, B.L.; Mashkovsky, I.S.; Markov, P.V.; Turova, O.V.; Tkachenko, O.P.; Pyryaev, P.A.; Bukhtiyarov, V.I. Liquid-phase hydrogenation of diphenylacetylene on Pd-Au/Al2O3 bimetallic catalysts. Russ. Chem. Bull. 2015, 64, 53–57. [Google Scholar] [CrossRef]
  211. He, J.-H.; Ichinose, I.; Kunitake, T.; Nakao, A.; Shiraishi, Y.; Toshima, N. Facile Fabrication of Ag−Pd Bimetallic Nanoparticles in ultrathin TiO2-gel films:  Nanoparticle morphology and catalytic activity. J. Am. Chem. Soc. 2003, 125, 11034–11040. [Google Scholar] [CrossRef]
  212. Toshima, N.; Yonezawa, T.; Kushihashi, K. Polymer-protected palladium–platinum bimetallic clusters: Preparation, catalytic properties and structural considerations. J. Chem. Soc. Faraday Trans. 1993, 89, 2537–2543. [Google Scholar] [CrossRef]
  213. Toshima, N.; Yonezawa, T. Bimetallic nanoparticles—Novel materials for chemical and physical applications. New J. Chem. 1998, 22, 1179–1201. [Google Scholar] [CrossRef]
  214. Kassube, J.K.; Gade, L.H. Immobilisation of the pyrphos ligand on soluble hyperbranched supports and use in rhodium-catalysed hydrogenation in ionic liquids. Adv. Synth. Catal. 2009, 351, 739–749. [Google Scholar] [CrossRef]
  215. Rodríguez, L.I.; Rossell, O.; Seco, M.; Muller, G. Carbosilane dendrimers peripherally functionalized with P-stereogenic diphosphine ligands and related rhodium complexes. J. Organomet. Chem. 2009, 694, 1938–1942. [Google Scholar] [CrossRef]
  216. Findeis, R.A.; Gade, L.H. Tripodal phosphane ligands with novel linker units and their rhodium complexes as building blocks for dendrimer catalysts. Eur. J. Inorg. Chem. 2003, 2003, 99–110. [Google Scholar] [CrossRef]
  217. Oosterom, G.E.; Steffens, S.; Reek, J.N.H.; Kamer, P.C.J.; van Leeuwen, P.W.N.M. Core-functionalized dendrimeric mono- and diphosphine rhodium complexes; Application in hydroformylation and hydrogenation. Top. Catal. 2002, 19, 61–73. [Google Scholar] [CrossRef]
  218. Zhang, F.; Li, Y.; Li, Z.-W.; He, Y.-M.; Zhu, S.-F.; Fan, Q.-H.; Zhou, Q.-L. Modular chiral dendritic monodentate phosphoramidite ligands for Rh(II)-catalyzed asymmetric hydrogenation: Unprecedented enhancement of enantioselectivity. Chem. Commun. 2008, 45, 6048–6050. [Google Scholar] [CrossRef] [PubMed]
  219. Nakamula, I.; Yamanoi, Y.; Yonezawa, T.; Imaoka, T.; Yamamoto, K.; Nishihara, H. Nanocage catalysts—Rhodium nanoclusters encapsulated with dendrimers as accessible and stable catalysts for olefin and nitroarene hydrogenations. Chem. Commun. 2008, 44, 5716–5718. [Google Scholar] [CrossRef] [PubMed]
  220. Menot, B.; Salmon, L.; Bouquillon, S. Platinum Nanoparticles Stabilized by Glycerodendrimers: Synthesis and Application to the Hydrogenation of α,β-Unsaturated Ketones under Mild Conditions. Eur. J. Inorg. Chem. 2015, 2015, 4518–4523. [Google Scholar] [CrossRef]
  221. Wang, Z.-J.; Deng, G.-J.; Li, Y.; He, Y.-M.; Tang, W.-J.; Fan, Q.-H. Enantioselective hydrogenation of quinolines catalyzed by Ir(BINAP)-cored dendrimers:  dramatic enhancement of catalytic activity. Org. Lett. 2007, 9, 1243–1246. [Google Scholar] [CrossRef] [PubMed]
  222. Wang, W.; Wang, Q. A fluorinated dendritic TsDPEN-Ru (II) catalyst for asymmetric transfer hydrogenation of prochiral ketones in aqueous media. Chem. Commun. 2010, 46, 4616–4618. [Google Scholar] [CrossRef]
  223. Maity, P.; Yamazoe, S.; Tsukuda, T. Dendrimer-encapsulated copper cluster as a chemoselective and regenerable hydrogenation catalyst. ACS Catal. 2013, 3, 182–185. [Google Scholar] [CrossRef]
  224. Johnson, J.A.; Makis, J.J.; Marvin, K.A.; Rodenbusch, S.E.; Stevenson, K.J. Size-dependent hydrogenation of p-Nitrophenol with pd nanoparticles synthesized with poly(amido)amine dendrimer templates. J. Phys. Chem. C 2013, 117, 44, 22644–22651. [Google Scholar] [CrossRef]
  225. Higaki, T.; Kitazawa, H.; Yamazoe, S.; Tsukuda, T. Partially oxidized iridium clusters within dendrimers: Size-controlled synthesis and selective hydrogenation of 2-nitrobenzaldehyde. Nanoscale 2016, 8, 11371–11374. [Google Scholar] [CrossRef] [Green Version]
  226. Bingwa, N.; Meijboom, R. Evaluation of catalytic activity of Ag and Au dendrimer-encapsulated nanoparticles in the reduction of 4-nitrophenol. J. Mol. Catal. A Chem. 2015, 396, 1–7. [Google Scholar] [CrossRef]
  227. Antonels, N.C.; Meijboom, R. Preparation of well-defined dendrimer encapsulated ruthenium nanoparticles and their evaluation in the reduction of 4-nitrophenol according to the langmuir–hinshelwood approach. Langmuir 2013, 29, 13433–13442. [Google Scholar] [CrossRef] [PubMed]
  228. Ma, Z.; Wu, R.; Han, Q.; Chen, R.; Gu, Z. Preparation of well-dispersed and anti-oxidized Ni nanoparticles using polyamidoamine dendrimers as templates and their catalytic activity in the hydrogenation of p-nitrophenol to p-aminophenol. Korean J. Chem. Eng. 2011, 28, 717–722. [Google Scholar] [CrossRef]
  229. Maeno, Z.; Mitsudome, T.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Selective synthesis of Rh5 carbonyl clusters within a polyamine dendrimer for chemoselective reduction of nitro aromatics. Chem. Commun. 2014, 50, 6526–6529. [Google Scholar] [CrossRef]
  230. Dumitrescu, I.; Crooks, R.M. Effect of mass transfer on the oxygen reduction reaction catalyzed by platinum dendrimer-encapsulated nanoparticles. Proc. Natl. Acad. Sci. USA 2012, 109, 11493–11497. [Google Scholar] [CrossRef] [Green Version]
  231. Mizugaki, T.; Miyauchi, Y.; Murata, M.; Ebitani, K.; Kaneda, K. Dendritic nanoreactor encapsulating Rh complex catalyst for hydroformylation. Chem. Lett. 2005, 34, 286–287. [Google Scholar] [CrossRef]
  232. Moore, E.; McInnes, S.J.; Vogt, A.; Voelcker, N.H. Rapid aqueous ‘click chemistry’ using Cu(I)-loaded dendrimers as macromolecular catalysts. Tetrahedron Lett. 2011, 52, 2327–2329. [Google Scholar] [CrossRef]
  233. Koishybay, A.; Shantz, D.F. Copper-gold nanoparticles encapsulated within surface-tethered dendrons as supported catalysts for the Click reaction. Appl. Catal. A Gen. 2018, 563, 196–203. [Google Scholar] [CrossRef]
  234. Mizugaki, T.; Murata, M.; Ooe, M.; Ebitani, K.; Kaneda, K. Novel catalysis of dendrimer-bound Pd (0) complexes: Sterically steered allylic amination and the first application for a thermomorphic system. Chem. Commun. 2002, 52–53. [Google Scholar] [CrossRef]
  235. Ooe, M.; Murata, M.; Takahama, A.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Self-assembled dendrimer-bound Pd (II) complexes via Acid-base interactions and their catalysis for allylic amination. Chem. Lett. 2003, 32, 692–693. [Google Scholar] [CrossRef]
  236. Ribaudo, F.; van Leeuwen, P.W.N.M.; Reek, J.N.H. Phosphorus functionalized dendrimers and hyperbranched polymers: Is there a need for perfect dendrimers in catalysis? Isr. J. Chem. 2009, 49, 79–98. [Google Scholar] [CrossRef]
  237. Gaab, M.; Bellemin-Laponnaz, S.; Gade, L.H. “Catalysis in a tea bag”: Synthesis, catalytic performance and recycling of dendrimer-immobilised bis- and trisoxazoline copper catalysts. Chem. Eur. J. 2009, 15, 5450–5462. [Google Scholar] [CrossRef] [PubMed]
  238. Laurent, R.; Caminade, A.M.; Majoral, J.P. A third generation chiral phosphorus-containing dendrimer as ligand in Pd-catalyzed asymmetric allylic alkylation. Tetrahedron Lett. 2005, 46, 6503–6506. [Google Scholar] [CrossRef]
  239. Köllner, C.; Togni, A. Synthesis, characterization, and application in asymmetric catalysis of dendrimers containing chiral ferrocenyl diphosphines. Can. J. Chem. 2001, 79, 1762–1774. [Google Scholar] [CrossRef]
  240. Keilitz, J.; Haag, R. Intramolecular Acceleration of asymmetric epoxide ring-opening by dendritic polyglycerol Salen–CrIII complexes. Eur. J. Org. Chem. 2009, 2009, 3272–3278. [Google Scholar] [CrossRef]
  241. Karakhanov, E.A.; Maximov, A.L.; Skorkin, V.A.; Zolotukhina, A.V.; Smerdov, A.S.; Tereshchenko, A.Y. Nanocatalysts based on dendrimers. Pure Appl. Chem. 2009, 81, 2013–2023. [Google Scholar] [CrossRef]
  242. Krishnan, G.R.; Sreekumar, K. Polystyrene-supported poly(amidoamine) dendrimer–manganese complex: Synthesis, characterization and catalysis. Appl. Catal. A Gen. 2009, 353, 80–86. [Google Scholar] [CrossRef]
  243. Jahier, C.; Sankar Mal, S.; Kortz, U.; Nlate, S. Dendritic zirconium-peroxotungstosilicate hybrids: Synthesis, characterization, and use as recoverable and reusable sulfide oxidation catalysts. Eur. J. Inorg. Chem. 2010, 2010, 1559–1566. [Google Scholar] [CrossRef]
  244. Iwasawa, T.; Tokunaga, M.; Obora, Y.; Tsuji, Y. Homogeneous palladium catalyst suppressing pd black formation in air oxidation of alcohols. J. Am. Chem. Soc. 2004, 126, 6554–6555. [Google Scholar] [CrossRef]
  245. Takahashi, M.; Imaoka, T.; Yamamoto, K. Reactivities of platinum subnanocluster catalysts for the oxidation reaction of alcohols. RSC Adv. 2015, 5, 100693–100696. [Google Scholar] [CrossRef]
  246. Sonobe, K.; Tanabe, M.; Yamamoto, K. Enhanced catalytic performance of subnano copper oxide particles. ACS Nano 2020, 14, 1804–1810. [Google Scholar] [CrossRef]
  247. Bhyrappa, P.; Young, J.K.; Moore, J.S.; Suslick, K.S. Shape selective epoxidation of alkenes by metalloporphyrin-dendrimers. J. Mol. Catal. A Chem. 1996, 113, 109–116. [Google Scholar] [CrossRef]
  248. Bhyrappa, P.; Young, J.; Moore, J.S.; Suslick, K.S. Dendrimer-metalloporphyrins: Synthesis and catalysis. J. Am. Chem. Soc. 1996, 118, 5708–5711. [Google Scholar] [CrossRef]
  249. Sellner, H.; Karjalainen, J.K.; Seebach, D. Preparation of dendritic and non-dendritic styryl-substituted salens for cross-linking suspension copolymerization with styrene and multiple use of the corresponding Mn and Cr complexes in enantioselective epoxidations and hetero-diels–alder reactions. Chem. Eur. J. 2001, 7, 2873–2887. [Google Scholar] [CrossRef]
  250. Plault, L.; Hauseler, A.; Nlate, S.; Astruc, D.; Ruiz, J.; Gatard, S.; Neumann, R. Synthesis of dendritic polyoxometalate complexes assembled by ionic bonding and their function as recoverable and reusable oxidation catalysts. Angew. Chem. Int. Ed. 2004, 43, 2924–2928. [Google Scholar] [CrossRef]
  251. Vasylyev, M.V.; Astruc, D.; Neumann, R. Dendritic phosphonates and the in situ assembly of polyperoxophosphotungstates: Synthesis and catalytic epoxidation of alkenes with hydrogen peroxide. Adv. Synth. Catal. 2005, 347, 39–44. [Google Scholar] [CrossRef]
  252. Bu, J.; Judeh, Z.M.A.; Ching, C.B.; Kawi, S. Epoxidation of olefins catalyzed by Mn (II) salen complex anchored on PAMAM–SiO2 dendrimer. Catal. Lett. 2003, 85, 183–187. [Google Scholar] [CrossRef]
  253. Peng, M.; Chen, Y.; Tan, R.; Zheng, W.; Yin, D. A highly efficient and recyclable catalyst: Dendrimer supported chiral salen Mn (III) complex for asymmetric epoxidation. RSC Adv. 2013, 3, 20684–20692. [Google Scholar] [CrossRef]
  254. Smith, G.; Chen, R.; Mapolie, S. The synthesis and catalytic activity of a first-generation poly(propylene imine) pyridylimine palladium metallodendrimer. J. Organomet. Chem. 2003, 673, 111–115. [Google Scholar] [CrossRef]
  255. Benito, J.M.; de Jesús, E.; de la Mata, F.J.; Flores, J.C.; Gómez, R. Neutral and cationic dendritic palladium (II) complexes containing N,N’-iminopyridine chelating ligands. Synthesis and their use for the syndiospecific copolymerization of CO/4-tert-butylstyrene. Organometallics 2006, 25, 3045–3055. [Google Scholar] [CrossRef]
  256. Müller, C.; Ackerman, L.J.; Reek, J.N.H.; Kamer, P.C.J.; van Leeuwen, P.W.N.M. Site-isolation effects in a dendritic nickel catalyst for the oligomerization of ethylene. J. Am. Chem. Soc. 2004, 126, 14960–14963. [Google Scholar] [CrossRef]
  257. Andres, R.; de Jesús, E.; de la Mata, F.J.; Flores, J.C.; Gómez, R. Titanocene and zirconocene complexes containing dendrimer-substituted cyclopentadienyl ligands—Synthesis and ethylene polymerization. Eur. J. Inorg. Chem. 2002, 2002, 2281–2286. [Google Scholar] [CrossRef]
  258. Mager, M.; Becke, S.; Windisch, H.; Denninger, U. Noncoordinating dendrimer polyanions: Cocatalysts for the metallocene-catalyzed olefin polymerization. Angew. Chem. Int. Ed. 2001, 40, 1898–1902. [Google Scholar] [CrossRef]
  259. Andres, R.; de Jesús, E.; de la Mata, F.J.; Flores, J.C.; Gómez, R. Dendritic-β-diketiminato titanium and zirconium complexes: Synthesis and ethylene polymerisation. J. Organometal. Chem. 2005, 690, 939–943. [Google Scholar] [CrossRef]
  260. Zheng, Z.-J.; Chen, J.; Li, Y.-S. The synthesis and catalytic of poly(bis(imino)pyridyl) iron (II) metallodendrimer. J. Organometal. Chem. 2004, 689, 3040–3045. [Google Scholar] [CrossRef]
  261. Benito, J.M.; de Jesús, E.; de la Mata, F.J.; Flores, J.C.; Gómez, R. Generation effects on the microstructure and product distribution in ethylene polymerization promoted by dendritic nickel catalysts. Chem. Commun. 2005, 41, 5217–5219. [Google Scholar] [CrossRef]
  262. Benito, J.M.; de Jesús, E.; de la Mata, F.J.; Flores, J.C.; Gómez, R.; Gómez-Sal, P. Mononuclear and dendritic nickel (II) complexes containing N,N′-iminopyridine chelating ligands:  Generation effects on the catalytic oligomerization and polymerization of ethylene. Organometallics 2006, 25, 3876–3887. [Google Scholar] [CrossRef]
  263. Malgas-Enus, R.; Mapolie, S.F.; Smith, G.S. Norbornene polymerization using multinuclear nickel catalysts based on a polypropyleneimine dendrimer scaffold. J. Organomet. Chem. 2008, 693, 2279–2286. [Google Scholar] [CrossRef]
  264. Gatard, S.; Nlate, S.; Cloutet, E.; Bravic, G.; Blais, J.-C.; Astruc, D. Dendritic stars by ring-opening-metathesis polymerization from ruthenium–carbene dendrimers. Angew. Chem. Int. Ed. 2003, 42, 452–456. [Google Scholar] [CrossRef]
  265. Gatard, S.; Kahlal, S.; Méry, D.; Nlate, S.; Cloutet, E.; Saillard, J.-Y.; Astruc, D. Synthesis, chemistry, DFT calculations and ROMP activity of monomeric benzylidene complexes containing a chelating diphosphine and of four generations of metallodendritic analogues. Positive and negative dendritic effects and formation of dendritic ruthenium polynorbornene stars. Organometallics 2004, 23, 1313–1324. [Google Scholar] [CrossRef]
  266. Méry, D.; Astruc, D. Synthesis of monomeric and dendritic ruthenium benzylidene cis–bis-tertiobutyl phosphine complexes that catalyze the ROMP of norbornene under ambient conditions. J. Mol. Catal. A Chem. 2005, 227, 1–5. [Google Scholar] [CrossRef]
  267. Garber, S.B.; Kingsbury, J.S.; Gray, B.L.; Hoveyda, A.H. Efficient and recyclable monomeric and dendritic Ru-based metathesis catalysts. J. Am. Chem. Soc. 2000, 122, 8168–8179. [Google Scholar] [CrossRef]
  268. Astruc, D.; Diallo, A.K.; Gatard, S.; Liang, L.; Ornelas, C.; Martinez, V.; Méry, D.; Ruiz, J. Olefin metathesis in nano-sized systems. Beilstein J. Org. Chem. 2011, 7, 94–103. [Google Scholar] [CrossRef] [Green Version]
  269. Diallo, A.K.; Boisselier, E.; Liang, L.; Ruiz, J.; Astruc, D. Dendrimer-induced molecular catalysis in water: The example of olefin metathesis. Chem. Eur. J. 2010, 16, 11832–11835. [Google Scholar] [CrossRef] [PubMed]
  270. Wijkens, P.; Jastrzebski, J.T.B.H.; van der Schaaf, P.A.; Kolly, R.; Hafner, A.; van Koten, G. Synthesis of periphery-functionalized dendritic molecules using polylithiated dendrimers as starting material. Org. Lett. 2000, 2, 1621–1624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  271. Maeno, Z.; Okao, M.; Mitsudome, T.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Regioselective oxidative coupling of 2,6-dimethylphenol to tetramethyldiphenoquinone using polyamine dendrimer-encapsulated Cu catalysts. RSC Adv. 2013, 3, 9662–9665. [Google Scholar] [CrossRef]
  272. Yamamoto, K.; Kawana, Y.; Tsuji, M.; Hayashi, M.; Imaoka, T. Additive-Free Synthesis of poly(phenylene oxide): Aerobic oxidative polymerization in a base-condensed dendrimer capsule. J. Am. Chem. Soc. 2007, 129, 9256–9257. [Google Scholar] [CrossRef] [PubMed]
  273. Imaoka, T.; Kawana, Y.; Tsuji, M.; Yamamoto, K. Modulation of the aerobic oxidative polymerization in phenylazomethine dendrimers assembling copper complexes. Chem. Eur. J. 2010, 16, 11003–11011. [Google Scholar] [CrossRef] [PubMed]
  274. Muraki, T.; Fujita, K.-I.; Kujime, M. Synthesis of novel dendritic 2,2’-bipyridine ligands and their application to lewis acid-catalyzed diels−alder and three-component condensation reactions. J. Org. Chem. 2007, 72, 7863–7870. [Google Scholar] [CrossRef] [PubMed]
  275. Garcia-Martinez, J.C.; Lezutekong, R.; Crooks, R.M. Dendrimer-encapsulated pd nanoparticles as aqueous, room-temperature catalysts for the stille reaction. J. Am. Chem. Soc. 2005, 127, 5097–5103. [Google Scholar] [CrossRef]
  276. Diallo, A.K.; Ornelas, C.; Salmon, L.; Ruiz Aranzaes, J.; Astruc, D. “Homeopathic” catalytic activity and atom-leaching mechanism in miyaura–suzuki reactions under ambient conditions with precise dendrimer-stabilized Pd nanoparticles. Angew. Chem. Int. Ed. 2007, 46, 8644–8648. [Google Scholar] [CrossRef]
  277. Yeung, L.K.; Lee, C.T., Jr.; Johnston, K.P.; Crooks, R.M. Catalysis in supercritical CO2 using dendrimer-encapsulated palladium nanoparticles. Chem. Commun 2001, 2290–2291. [Google Scholar] [CrossRef]
  278. Lemo, J.; Heuzé, K.; Astruc, D. Synthesis and catalytic activity of DAB-dendrimer encapsulated Pd nanoparticles for the Suzuki coupling reaction. Inorg. Chim. Acta 2006, 359, 4909–4911. [Google Scholar] [CrossRef]
  279. Enomoto, M.; Aida, T. Iron complexes capped with dendrimer-appended triazacyclononanes as the novel spatially encumbered models of non-heme iron proteins. J. Am. Chem. Soc. 2002, 124, 6099–6108. [Google Scholar] [CrossRef]
  280. Enomoto, M.; Aida, T. Self-Assembly of a copper-ligating dendrimer that provides a new non-heme metalloprotein mimic: “Dendrimer effects” on stability of the bis(µ-oxo)dicopper (III) core. J. Am. Chem. Soc. 1999, 121, 874–875. [Google Scholar] [CrossRef]
  281. Helms, B.; Fréchet, J.M.J. The dendrimer effect in homogeneous catalysis. Adv. Synth. Catal. 2006, 348, 1125–1148. [Google Scholar] [CrossRef]
  282. Caminade, A.M.; Ouali, A.; Laurent, R.; Turrin, C.O.; Majoral, J.P. The dendritic effect illustrated with phosphorus dendrimers. Chem. Soc. Rev. 2015, 44, 3890–3899. [Google Scholar] [CrossRef]
  283. Mansour, A.; Kehat, T.; Portnoy, M. Dendritic effects in catalysis by Pd complexes of bidentate phosphines on a dendronized support: Heck vs. carbonylation reactions. Org. Biomol. Chem. 2008, 6, 3382–3387. [Google Scholar] [CrossRef]
  284. Breinbauer, R.; Jacobsen, E.N. Cooperative asymmetric catalysis with dendrimeric [Co(salen)] complexes. Angew. Chem. Int. Ed. 2000, 39, 3604–3607. [Google Scholar] [CrossRef]
  285. Boitiaux, J.P.; Cosyns, J.; Vasijdevan, S. Hydrogenation of highly unsaturated hydrocarbons over highly dispersed Pd catalyst. Part II: Ligand effect of piperidine. Appl. Catal. 1985, 15, 317–326. [Google Scholar] [CrossRef]
  286. Aramendia, M.A.; Borau, V.; Jimenez, C.; Marinas, J.M.; Sempere, M.E.; Urbano, F.J. Optimization of the selective semi-hydrogenation of phenylacetylene with supported palladium systems. Appl. Catal. 1990, 63, 375–389. [Google Scholar] [CrossRef]
  287. Karakhanov, E.A.; Maximov, A.L.; Zolotukhina, A.V. Selective semi-hydrogenation of phenyl acetylene by Pd nanocatalysts encapsulated into dendrimer networks. Mol. Catal. 2019, 469, 98–110. [Google Scholar] [CrossRef]
  288. Bond, G.C.; Rawle, A.F. Catalytic hydrogenation in the liquid phase. Part 1. Hydrogenation of isoprene catalyzed by palladium, palladium−gold, and palladium−silver catalysts. J. Mol. Catal. A Chem. 1996, 109, 261–271. [Google Scholar] [CrossRef]
  289. Bond, G.C. Product selectivities in isoprene hydrogenation: Diagnosis of π-allylic intermediates. J. Mol. Catal. A Chem. 1997, 118, 333–339. [Google Scholar] [CrossRef]
  290. Hoover, N.N.; Auten, B.J.; Chandler, B.D. Tuning supported catalyst reactivity with dendrimer-templated Pt−Cu nanoparticles. J. Phys. Chem. B. 2006, 110, 8606–8612. [Google Scholar] [CrossRef] [Green Version]
  291. Lang, H.; May, R.A.; Iversen, B.L.; Chandler, B.D. Dendrimer-encapsulated nanoparticle precursors to supported platinum catalysts. J. Am. Chem. Soc. 2003, 125, 14832–14836. [Google Scholar] [CrossRef] [Green Version]
  292. Huang, W.; Kuhn, J.N.; Tsung, C.-K.; Zhang, Y.; Habas, S.E.; Yang, P.; Somorjai, G.A. Dendrimer templated synthesis of one nanometer Rh and Pt particles supported on mesoporous silica: Catalytic activity for ethylene and pyrrole hydrogenation. Nano Lett. 2008, 8, 2027–2034. [Google Scholar] [CrossRef] [Green Version]
  293. Ye, R.; Zhukhovitskly, A.V.; Deraedt, C.; Toste, F.D.; Somorjai, G.A. Supported dendrimer-encapsulated metal clusters: Toward heterogenizing homogeneous catalysts. Acc. Chem. Res. 2017, 50, 1894–1901. [Google Scholar] [CrossRef] [Green Version]
  294. Singh, A.; Chandler, B.D. Low-temperature activation conditions for PAMAM dendrimer templated Pt nanoparticles. Langmuir 2005, 21, 10776–10782. [Google Scholar] [CrossRef] [Green Version]
  295. Rogers, M.C.; Adisa, B.; Bruce, D.A. Synthesis and Characterization of dendrimer-templated mesoporous oxidation catalysts. Catal. Lett. 2004, 98, 29–36. [Google Scholar] [CrossRef]
  296. Choi, H.C.; Kim, W.; Wang, D.; Dai, H. Delivery of catalytic metal species onto surfaces with dendrimer carriers for the synthesis of carbon nanotubes with narrow diameter distribution. J. Phys. Chem. B. 2002, 106, 12361–12365. [Google Scholar] [CrossRef]
  297. Beakley, L.W.; Yost, S.E.; Cheng, R.; Chandler, B.D. Nanocomposite catalysts: Dendrimer encapsulated nanoparticles immobilized in sol–gel silica. Appl. Catal. A Gen. 2005, 292, 124–129. [Google Scholar] [CrossRef]
  298. Kumar, P.A.; Ha, H.P. Synthesis and dispersion of dendrimer-encapsulated Pt nanoparticles on γ-Al2O3 for the reduction of NOx by methane. Catal. Lett. 2010, 136, 177–184. [Google Scholar] [CrossRef]
  299. Liu, D.X.; Lopez de Jesus, Y.M.; Monnier, J.R.; Williams, C.T. Preparation, characterization, and kinetic evaluation of dendrimer-derived bimetallic Pt–Ru/SiO2 catalysts. J. Catal. 2010, 269, 376–387. [Google Scholar] [CrossRef]
  300. Lafaye, G.; Williams, C.T.; Amiridis, M.D. Synthesis and microscopic characterization of dendrimer-derived Ru/Al2O3 catalysts. Catal. Lett. 2004, 96, 43–47. [Google Scholar] [CrossRef]
  301. Lafaye, G.; Siani, A.; Marécot, P.; Amiridis, M.D.; Williams, C.T. Particle size control in dendrimer-derived supported ruthenium catalysts. J. Phys. Chem. B 2006, 110, 7725–7731. [Google Scholar] [CrossRef]
  302. Kuhn, J.N.; Huang, W.; Tsung, C.-K.; Zhang, Y.; Somorjaim, G.A. Structure Sensitivity of carbon−nitrogen ring opening: Impact of platinum particle size from below 1 to 5 nm upon pyrrole hydrogenation product selectivity over monodisperse platinum nanoparticles loaded onto mesoporous silica. J. Am. Chem. Soc. 2008, 130, 14026–14027. [Google Scholar] [CrossRef] [Green Version]
  303. Deraedt, C.; Ye, R.; Ralston, W.T.; Toste, F.D.; Somorjai, G.A. Dendrimer-Stabilized metal nanoparticles as efficient catalysts for reversible dehydrogenation/hydrogenation of n-heterocycles. J. Am. Chem. Soc. 2017, 139, 18084–18092. [Google Scholar] [CrossRef] [Green Version]
  304. Deraedt, C.; Melaet, G.; Ralston, W.T.; Ye, R.; Somorjai, G.A. Platinum and other transition metal nanoclusters (Pd, Rh) stabilized by pamam dendrimer as excellent heterogeneous catalysts: Application to the methylcyclopentane (MCP) Hydrogenative isomerization. Nano Lett. 2017, 17, 1853–1862. [Google Scholar] [CrossRef]
  305. Ye, R.; Yuan, B.; Zhao, J.; Ralston, W.T.; Wu, C.-Y.; Barin, E.U.; Toste, D.F.; Somorjai, G.A. Metal Nanoparticles catalyzed selective carbon–carbon bond activation in the liquid phase. J. Am. Chem. Soc. 2016, 138, 8533–8537. [Google Scholar] [CrossRef]
  306. Witham, C.A.; Huang, W.; Tsung, C.-K.; Kuhn, J.N.; Somorjai, G.A.; Toste, F.D. Converting homogeneous to heterogeneous in electrophilic catalysis using monodisperse metal nanoparticles. Nat. Chem. 2010, 2, 36–41. [Google Scholar] [CrossRef]
  307. Huang, W.; Liu, J.H.-C.; Alayoglu, P.; Li, Y.; Witham, C.A.; Tsung, C.-K.; Toste, F.D.; Somorjai, G.A. highly active heterogeneous palladium nanoparticle catalysts for homogeneous electrophilic reactions in solution and the utilization of a continuous flow reactor. J. Am. Chem. Soc. 2010, 132, 16771–16773. [Google Scholar] [CrossRef]
  308. Li, Y.; Liu, J.H.-C.; Witham, C.A.; Huang, W.; Marcus, M.A.; Fakra, S.C.; Alayoglu, P.; Zhu, Z.; Thompson, C.M.; Arjun, A.; et al. A Pt-cluster-based heterogeneous catalyst for homogeneous catalytic reactions: X-ray absorption spectroscopy and reaction kinetic studies of their activity and stability against leaching. J. Am. Chem. Soc. 2011, 133, 13527–13533. [Google Scholar] [CrossRef]
  309. Gross, E.; Shu, X.-Z.; Alayoglu, S.; Bechtel, H.A.; Martin, M.C.; Toste, F.D.; Somorjai, G.A. In situ IR and X-ray high spatial-resolution microspectroscopy measurements of multistep organic transformation in flow microreactor catalyzed by au nanoclusters. J. Am. Chem. Soc. 2014, 136, 3624–3629. [Google Scholar] [CrossRef]
  310. Gross, E.; Lui, J.H.-C.; Toste, F.D.; Somorjai, G.A. Control of selectivity in heterogeneous catalysis by tuning nanoparticle properties and reactor residence time. Nat. Chem. 2012, 4, 947–952. [Google Scholar] [CrossRef]
  311. Ye, R.; Zhao, J.; Yuan, B.; Liu, W.-C.; De Araujo, J.R.; Faucher, F.F.; Chang, M.; Deraedt, C.V.; Toste, F.D.; Somorjai, G.A. New insights into aldol reactions of methyl isocyanoacetate catalyzed by heterogenized homogeneous catalysts. Nano Lett. 2017, 17, 584–589. [Google Scholar] [CrossRef]
  312. Antonels, N.C.; Meijboom, R. Preparation of well-defined dendrimer encapsulated ruthenium nanoparticles and their application as catalyst and enhancement of activity when utilised as SCILL catalysts in the hydrogenation of citral. Catal. Commun. 2014, 57, 148–152. [Google Scholar] [CrossRef]
  313. Antonels, N.C.; Williams, M.B.; Meijboom, R.; Haumann, M. Well-defined dendrimer encapsulated ruthenium SCILL catalysts for partial hydrogenation of toluene in liquid-phase. J. Mol. Catal. A Chem. 2016, 421, 156–160. [Google Scholar] [CrossRef]
  314. Zhao, M.; Liu, Y.; Crooks, R.M.; Bergbreiter, D.E. Preparation of highly impermeable hyperbranched-polymer thin film coatings using dendrimers first as building blocks and then as in-situ thermosetting agents. J. Am. Chem. Soc. 1999, 121, 923–930. [Google Scholar] [CrossRef]
  315. Foppa, L.; Dupont, J. Benzene partial hydrogenation: Advances and perspectives. Chem. Soc. Rev. 2015, 44, 1886–1897. [Google Scholar] [CrossRef]
  316. Satagopan, V.; Chandalia, S.B. Selectivity aspects in the multi-phase hydrogenation of α, β-unsaturated aldehydes over supported noble metal catalysts: Part I. J. Chem. Technol. Biotechnol. 1994, 59, 257–263. [Google Scholar] [CrossRef]
  317. Maiti, P.K.; Çaǧın, T.; Lin, S.h.-T.; Goddard, W.A. Effect of solvent and pH on the structure of PAMAM dendrimers. Macromolecules 2005, 38, 979–991. [Google Scholar] [CrossRef]
  318. Gosika, M.; Maiti, P.K. pH and generation dependent morphologies of PAMAM dendrimers on a graphene substrate. Soft Matter 2018, 14, 1925–1938. [Google Scholar] [CrossRef] [PubMed]
  319. Koper, G.J.M.; van Genderen, M.H.P.; Elissen-Román, C.; Baars, M.W.P.L.; Meijer, E.W.; Borkovec, M. Protonation mechanism of poly(propylene imine) dendrimers and some associated oligo amines. J. Am. Chem. Soc. 1997, 119, 6512–6521. [Google Scholar] [CrossRef] [Green Version]
  320. Soeiro Suppino, R.; Landers, R.; Gomez Cobo, A.J. Partial hydrogenation of benzene on Ru catalysts: Effects of additives in the reaction medium. Appl. Catal. A Gen. 2013, 452, 9–16. [Google Scholar] [CrossRef]
  321. Liu, J.-L.; Zhu, L.-J.; Pei, Y.; Zhuang, J.-H.; Li, H.; Li, H.-X.; Qiao, M.-H.; Fan, K.-N. Ce-promoted Ru/SBA-15 catalysts prepared by a “two solvents” impregnation method for selective hydrogenation of benzene to cyclohexene. Appl. Catal. A Gen. 2009, 353, 282–287. [Google Scholar] [CrossRef]
  322. Sun, H.; Pan, Y.; Li, S.; Zhang, Y.; Dong, Y.; Liu, S.; Liu, Z. Selective hydrogenation of benzene to cyclohexene over Ce-promoted Ru catalysts. J. Energy Chem. 2013, 22, 710–716. [Google Scholar] [CrossRef]
  323. Hu, S.-C.; Chen, Y.-W. Partial hydrogenation of benzene on Ru−Zn/SiO2 Catalysts. Ind. Eng. Chem. Res. 2001, 40, 6099–6104. [Google Scholar] [CrossRef]
  324. Karakhanov, E.A.; Maksimov, A.L.; Zolotukhina, A.V.; Vinokurov, V.A. Oxidation of p-Xylene. A review. Russ. J. Appl. Chem. 2018, 91, 707–727. [Google Scholar] [CrossRef]
  325. Karakhanov, E.; Maximov, A.; Zolotukhina, A.; Vinokurov, V.; Ivanov, E.; Glotov, A. Manganese and cobalt doped hierarchical mesoporous halloysite-based catalysts for selective oxidation of p-Xylene to terephthalic. Acid. Catal. 2019, 10, 7. [Google Scholar] [CrossRef] [Green Version]
  326. Kim, N.R.; Shin, K.; Jung, I.; Shim, M.; Lee, H.M. Ag−Cu bimetallic nanoparticles with enhanced resistance to oxidation: A combined experimental and theoretical study. J. Phys. Chem. C 2014, 118, 26324–26331. [Google Scholar] [CrossRef]
  327. Weihua, W.; Xuelin, T.; Kai, C.; Gengyu, C. Synthesis and characterization of Pt–Cu bimetallic alloy nanoparticles by reverse micelles method. Colloids Surf. A Physicochem. Eng. Asp. 2006, 273, 35–42. [Google Scholar] [CrossRef]
  328. Ye, H.; Crooks, R.M. Effect of elemental composition of PtPd bimetallic nanoparticles containing an average of 180 atoms on the kinetics of the electrochemical oxygen reduction reaction. J. Am. Chem. Soc. 2007, 129, 3627–3633. [Google Scholar] [CrossRef]
  329. Vijayaraghavan, G.; Stevenson, K.J. Synergistic Assembly of dendrimer-templated platinum catalysts on nitrogen-doped carbon nanotube electrodes for oxygen reduction. Langmuir 2007, 23, 5279–5282. [Google Scholar] [CrossRef]
  330. Frasconi, M.; Tortolini, C.; Botre, F.; Mazzei, F. Multifunctional Au nanoparticle dendrimer-based surface plasmon resonance biosensor and its application for improved insulin detection. Anal. Chem. 2010, 82, 7335–7342. [Google Scholar] [CrossRef]
  331. Ghiaci, M.; Mostajeran, M.; Gil, A. Synthesis and characterization of Co−Mn Nanoparticles immobilized on a modified bentonite and its application for oxidation of p-Xylene to terephthalic. Acid. Ind. Eng. Chem. Res. 2012, 51, 15821–15831. [Google Scholar] [CrossRef]
  332. Antebi, S.; Arya, P.; Manzer, L.E.; Alper, H. Carbonylation reactions of iodoarenes with PAMAM dendrimer-palladium catalysts immobilized on silica. J. Org. Chem. 2002, 67, 6623–6631. [Google Scholar] [CrossRef] [Green Version]
  333. Wu, H.; Liu, Z.; Wang, X.; Zhao, B.; Zhang, J.; Li, C. Preparation of hollow capsule-stabilized gold nanoparticles through the encapsulation of the dendrimer. J. Colloid Interface Sci. 2006, 302, 142–148. [Google Scholar] [CrossRef]
  334. Chanthateyanonth, R.; Alper, H. Recyclable Tridentate stable palladium (II) PCP-type catalysts supported on silica for the selective synthesis of lactones. Adv. Synth. Catal. 2004, 346, 1375–1384. [Google Scholar] [CrossRef]
  335. Bourque, S.C.; Alper, H.; Manzer, L.E.; Arya, P. Hydroformylation reactions using recyclable rhodium-complexed dendrimers on silica. J. Am. Chem. Soc. 2000, 122, 956–957. [Google Scholar] [CrossRef]
  336. Touzani, R.; Alper, H. PAMAM dendrimer-palladium complex catalyzed synthesis of five, six or seven-membered ring lactones and lactams by cyclocarbonylation methodology. J. Mol. Catal. A Chem. 2005, 227, 197–207. [Google Scholar] [CrossRef]
  337. Lu, S.-M.; Alper, H. Intramolecular carbonylation reactions with recyclable palladium-complexed dendrimers on silica:  Synthesis of oxygen, nitrogen, or sulfur-containing medium ring fused heterocycles. J. Am. Chem. Soc. 2005, 127, 14776–14784. [Google Scholar] [CrossRef] [PubMed]
  338. Alper, H.; Arya, P.; Bourque, S.C.; Jefferson, G.R.; Manzer, L.E. Heck reaction using palladium complexed to dendrimers on silica. Can. J. Chem. 2000, 78, 920–924. [Google Scholar] [CrossRef]
  339. Tsubokawa, N.; Ichioka, H.; Satoh, T.; Hayashi, S.; Fujiki, K. Grafting of dendrimer-like highly branched polymer onto Ultrafine silica surface. React. Funct. Polym. 1998, 37, 75–82. [Google Scholar] [CrossRef]
  340. Chanthateyanonth, R.; Alper, H. The first synthesis of stable palladium (II) PCP-type catalysts supported on silica—Application to the heck reaction. J. Mol. Catal. A Chem. 2003, 201, 23–31. [Google Scholar] [CrossRef]
  341. Xu, Y.; Zhang, Z.; Zheng, J.; Du, Q.; Li, Y. Synthesis of dendrimers terminated by DABCO ligands and applications of its palladium nanoparticles for catalyzing Suzuki–Miyaura and Mizoroki–Heck couplings. Appl. Organometal. Chem. 2013, 27, 13–18. [Google Scholar] [CrossRef]
  342. Landarani Isfahani, A.; Mohammadpoor-Baltork, I.; Mirkhani, V.; Khosropour, A.R.; Moghadam, M.; Tangestaninejad, S.; Kia, R. Palladium nanoparticles immobilized on nano-silica triazine dendritic polymer (Pdnp-nSTDP): An efficient and reusable catalyst for suzuki–miyaura cross-coupling and heck reactions. Adv. Synth. Catal. 2013, 355, 957–972. [Google Scholar] [CrossRef] [Green Version]
  343. Hagiwara, H.; Sasaki, H.; Tsubokawa, N.; Hoshi, T.; Suzuki, T.; Tsuda, T.; Kuwabata, S. Immobilization of Pd on nanosilica dendrimer as SILC: Highly active and sustainable cluster catalyst for suzuki-miyaura reaction. Synlett 2010, 2010, 1990–1996. [Google Scholar] [CrossRef]
  344. Ohtaka, A.; Kawase, M.; Usami, A.; Fukui, S.; Yamashita, M.; Yamaguchi, K.; Sakon, A.; Shiraki, T.; Ishida, T.; Nagata, S. Mechanistic study on allylic arylation in water with linear polystyrene-stabilized Pd and PdO nanoparticles. ACS Omega 2019, 4, 15764–15770. [Google Scholar] [CrossRef] [Green Version]
  345. Landarani-Isfahani, A.; Mohammadpoor-Baltork, I.; Mirkhani, V.; Moghadam, M.; Tangestaninejad, S.; Amiri Rudbari, H. Palladium nanoparticles immobilized on a nanosilica triazine dendritic polymer: A recyclable and sustainable nanoreactor for C–S cross-coupling. RSC Adv. 2020, 10, 21198–21205. [Google Scholar] [CrossRef]
  346. Eremin, D.B.; Ananikov, V.P. Understanding active species in catalytic transformations: From molecular catalysis to nanoparticles, leaching, “Cocktails” of catalysts and dynamic systems. Coord. Chem. Rev. 2017, 346, 2–19. [Google Scholar] [CrossRef]
  347. Polynski, M.V.; Ananikov, V.P. Modeling key pathways proposed for the formation and evolution of “Cocktail”-type systems in Pd-catalyzed reactions involving ArX reagents. ACS Catal. 2019, 9, 3991–4005. [Google Scholar] [CrossRef] [Green Version]
  348. Galushko, A.S.; Prima, D.O.; Burykina, J.V.; Ananikov, V.P. Comparative study of aryl halides in Pd-mediated reactions: Key factors beyond the oxidative addition step. Inorg. Chem. Front. 2021, 8, 620–635. [Google Scholar] [CrossRef]
  349. Yoshikawa, S.; Satoh, T.; Tsubokawa, N. Post-grafting of polymer with controlled molecular weight onto silica surface by termination of living polymer cation with terminal amino groups of dendrimer-grafted ultrafine silica. Colloids Surfaces A Physicochem. Eng. Asp. 1999, 153, 395–399. [Google Scholar] [CrossRef]
  350. Murota, M.; Sato, S.; Tsubokawa, N. Scale-up synthesis of hyperbranched Poly(amidoamine)-grafted ultrafine silica using dendrimer synthesis methodology in solvent-free dry-system. Polym. Adv. Technol. 2002, 13, 144–150. [Google Scholar] [CrossRef]
  351. Reynhardt, J.P.K.; Yang, Y.; Sayari, A.; Alper, H. Polyamidoamine dendrimers prepared inside the channels of pore-expanded periodic mesoporous silica. Adv. Funct. Mater. 2005, 15, 1641–1646. [Google Scholar] [CrossRef]
  352. Wang, C.L.; Zhu, G.S.; Li, J.; Cai, X.H.; Wei, Y.H.; Zhang, D.L.; Qiu, S.L. Rigid nanoscopic containers for highly dispersed, stable metal and bimetal nanoparticles with both size and site control. Chem. Eur. J. 2005, 11, 4975–4982. [Google Scholar] [CrossRef]
  353. Jiang, Y.; Gao, Q. Heterogeneous hydrogenation catalyses over recyclable Pd (0) nanoparticle catalysts stabilized by PAMAM-SBA-15 organic−inorganic hybrid composites. J. Am. Chem. Soc. 2006, 128, 716–717. [Google Scholar] [CrossRef]
  354. Li, H.; Lü, J.; Zheng, Z.; Cao, R. An efficient and reusable silica/dendrimer supported platinum catalyst for electron transfer reactions. J. Colloid Interface Sci. 2011, 353, 149–155. [Google Scholar] [CrossRef]
  355. Li, H.; Zheng, Z.; Cao, M.; Cao, R. Stable gold nanoparticle encapsulated in silica-dendrimers organic−inorganic hybrid composite as recyclable catalyst for oxidation of alcohol. Micropor. Mesopor. Mater. 2010, 136, 42–49. [Google Scholar] [CrossRef]
  356. Niu, Y.; Yeung, L.K.; Crooks, R.M. Size-selective hydrogenation of olefins by dendrimer-encapsulated palladium nanoparticles. J. Am. Chem. Soc. 2001, 123, 6840–6846. [Google Scholar] [CrossRef]
  357. Zheng, Z.; Li, H.; Liu, T.; Cao, R. Monodisperse noble metal nanoparticles stabilized in SBA-15: Synthesis, characterization and application in microwave-assisted Suzuki–Miyaura coupling reaction. J. Catal. 2010, 270, 268–274. [Google Scholar] [CrossRef]
  358. Suijkerbuijk, B.M.J.M.; Shu, L.J.; Gebbink, R.J.M.K.; Schlüter, A.D.; van Koten, G. Single-site catalysts on a cylindrical support beyond nanosize. Organometallics 2003, 22, 4175–4177. [Google Scholar] [CrossRef] [Green Version]
  359. Schlüter, A.-D. Dendrimers with Polymeric Core: Towards Nanocylinders. In Dendrimers. Topics in Current Chemistry; de Meijere, A., Ed.; Springer: Berlin/Heidelberg, Germany, 1998; Volume 197, pp. 165–191. [Google Scholar] [CrossRef]
  360. Schlüter, A.D.; Rabe, J.P. Dendronized Polymers: Synthesis, characterization, assembly at interfaces, and manipulation. Angew. Chem. Int. Ed. 2000, 39, 864–883. [Google Scholar] [CrossRef]
  361. Frauenrath, H. Dendronized polymers—Building a new bridge from molecules to nanoscopic objects. Prog. Polym. Sci. 2005, 30, 325–384. [Google Scholar] [CrossRef]
  362. Boronoev, M.P.; Subbotina, E.S.; Kurmaeva, A.A.; Kardasheva, Y.S.; Maximov, A.L.; Karakhanov, E.A. Platinum and palladium nanoparticles in modified mesoporous phenol–formaldehyde polymers as hydrogenation catalysts. Petrol. Chem. 2016, 56, 109–120. [Google Scholar] [CrossRef]
  363. Kulikov, L.; Kalinina, M.; Makeeva, D.; Maximov, A.; Kardasheva, Y.; Terenina, M.; Karakhanov, E. Palladium catalysts based on porous aromatic frameworks, modified with ethanolamino-groups, for hydrogenation of alkynes, alkenes and dienes. Catalysts 2020, 10, 1106. [Google Scholar] [CrossRef]
  364. Krishnan, G.R.; Sreekumar, K. Characterization of polystyrene supported catalytically active poly(amidoamine) dendrimer-palladium nanoparticle conjugates. Soft Mater. 2010, 8, 114–129. [Google Scholar] [CrossRef]
  365. Sparrow, J.T. An improved polystyrene support for solid phase peptide synthesis. J. Org. Chem. 1976, 41, 1350–1353. [Google Scholar] [CrossRef]
  366. Krishnan, G.R.; Sreekumar, K. First example of organocatalysis by polystyrene-supported PAMAM dendrimers: Highly efficient and reusable catalyst for knoevenagel condensations. Eur. J. Org. Chem. 2008, 2008, 4763–4768. [Google Scholar] [CrossRef]
  367. Zheng, J.; Lin, S.; Zhu, X.; Jiang, B.; Yang, Z.; Pan, Z. Reductant directed formation of PS−PAMAM-supported gold nanoparticles for use as highly active and recyclable catalysts for the aerobic oxidation of alcohols and the homocoupling of phenylboronic acids. Chem. Commun. 2012, 48, 6235–6237. [Google Scholar] [CrossRef]
  368. Guo, Z.; Feng, H.; Ma, H.-C.; Kang, Q.-X.; Yang, Z.-W. Hydrogenation of five-membered heterocycles of polymer-supported palladium catalyst at normal temperature and pressure. Polym. Adv. Technol. 2004, 15, 100–104. [Google Scholar] [CrossRef]
  369. Arya, P.; Rao, N.V.; Singkhonrat, J.; Alper, H.; Bourque, S.C.; Manzer, L.E. A divergent, solid-phase approach to dendritic ligands on beads. heterogeneous catalysis for hydroformylation reactions. J. Org. Chem. 2000, 65, 1881–1885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  370. Dahan, A.; Portnoy, M. Remarkable dendritic effect in the polymer-supported catalysis of the heck arylation of olefins. Org. Lett. 2003, 5, 1197–1200. [Google Scholar] [CrossRef] [PubMed]
  371. Dahan, A.; Portnoy, M. Synthesis of Poly(aryl benzyl ether) dendrimers on solid support. Macromolecules 2003, 36, 1034–1038. [Google Scholar] [CrossRef]
  372. Reza Nabid, M.; Bide, Y.; Tabatabaei Rezaei, S.J. Pd nanoparticles immobilized on PAMAM-grafted MWCNTs hybrid materials as new recyclable catalyst for Mizoraki–Heck cross-coupling reactions. Appl. Catal. A Gen. 2011, 406, 124–132. [Google Scholar] [CrossRef]
  373. Basso, A.; Evans, B.; Pegg, N.; Bradley, M. Solid phase synthesis of aryl-ether dendrimers. Chem. Commun. 2001, 8, 697–698. [Google Scholar] [CrossRef]
  374. Dahan, A.; Portnoy, M. Dendritic effect in polymer-supported catalysis of the intramolecular Pauson–Khand reaction. Chem. Commun. 2002, 22, 2700–2701. [Google Scholar] [CrossRef]
  375. Ben, T.; Ren, H.; Shengqian, M.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.; Simmons, J.M.; et al. Targeted synthesis of a porous aromatic framework with high stability and exceptionally high surface area. Angew. Chem. Int. Ed. 2009, 48, 9457–9460. [Google Scholar] [CrossRef]
  376. Yuan, Y.; Sun, F.; Ren, H.; Jing, X.; Wang, W.; Ma, H.; Zhao, H.; Zhu, G. Targeted synthesis of a porous aromatic framework with a high adsorption capacity for organic molecules. J. Mater. Chem. 2011, 21, 13498–13502. [Google Scholar] [CrossRef]
  377. Maximov, A.; Zolotukhina, A.; Kulikov, L.; Kardasheva, Y.; Karakhanov, E. Ruthenium catalysts based on mesoporous aromatic frameworks for the hydrogenation of arenes. React. Kinet. Mech. Catal. 2016, 117, 729–743. [Google Scholar] [CrossRef]
  378. Karakhanov, E.; Kardasheva, Y.; Kulikov, L.; Maximov, A.; Zolotukhina, A.; Vinnikova, M.; Ivanov, A. Sulfide catalysts supported on porous aromatic frameworks for naphthalene hydroprocessing. Catalysts 2016, 6, 122. [Google Scholar] [CrossRef] [Green Version]
  379. Karakhanov, E.; Maximov, A.; Terenina, M.; Vinokurov, V.; Kulikov, L.; Makeeva, D.; Glotov, A. Selective hydrogenation of terminal alkynes over palladium nanoparticles within the pores of amino-modified porous aromatic frameworks. Catal. Today 2020, 357, 176–184. [Google Scholar] [CrossRef]
  380. Syrgiannis, Z.; La Parola, V.; Hadad, C.; Lucío, M.; Vázquez, E.; Giacalone, F.; Prato, M. An Atom-economical approach to functionalized single-walled carbon nanotubes: Reaction with disulfides. Angew. Chem. Int. Ed. 2013, 52, 6480–6483. [Google Scholar] [CrossRef]
  381. Tsubokawa, N.; Takayama, T. Surface modification of chitosan powder by grafting of dendrimer-like hyperbranched polymer onto the surface. React. Funct. Polym. 2000, 43, 341–350. [Google Scholar] [CrossRef]
  382. Giacalone, F.; Campisciano, V.; Calabrese, C.; La Parola, V.; Syrgiannis, Z.; Prato, M.; Gruttadauria, M. Single-Walled carbon nanotube–polyamidoamine dendrimer hybrids for heterogeneous catalysis. ACS Nano 2016, 10, 4627–4636. [Google Scholar] [CrossRef]
  383. Phan, N.T.S.; Van Der Sluys, M.; Jones, C.W. On the nature of the active species in palladium catalyzed mizoroki–heck and suzuki–miyaura couplings—Homogeneous or heterogeneous catalysis. A critical review. Adv. Synth. Catal. 2006, 348, 609–679. [Google Scholar] [CrossRef]
  384. Thathagar, M.B.; ten Elshof, J.E.; Rothenberg, G. Pd Nanoclusters in C–C coupling reactions: Proof of leaching. Angew. Chem. Int. Ed. 2006, 45, 2886–2890. [Google Scholar] [CrossRef]
  385. Jiang, G.; Wang, L.; Chen, C.; Dong, X.; Chen, T.; Yu, H. Study on attachment of highly branched molecules onto multiwalled carbon nanotubes. Mater Lett. 2005, 59, 2085–2089. [Google Scholar] [CrossRef]
  386. Hwang, S.-H.; Moorefield, C.N.; Wang, P.; Jeong, K.-U.; Cheng, S.Z.D.; Kotta, K.K.; Newkome, G.R. Dendron-tethered and templated CdS quantum dots on single-walled carbon nanotubes. J. Am. Chem. Soc. 2006, 128, 23–7505. [Google Scholar] [CrossRef]
  387. Yamamoto, D.; Koshiyama, T.; Watanabe, S.; Miyahara, M.T. Synthesis and photoluminescence characterization of dendrimer-encapsulated CdS quantum dots. Colloids Surf. A Physicochem. Eng. Asp. 2012, 411, 12–17. [Google Scholar] [CrossRef] [Green Version]
  388. Hwang, S.-H.; Moorefield, C.N.; Wang, P.; Jeong, K.-U.; Cheng, S.Z.D.; Kotta, K.K.; Newkome, G.R. Construction of CdS quantum dots via a regioselective dendritic functionalized cellulose template. Chem. Commun. 2006, 33, 3495–3497. [Google Scholar] [CrossRef]
  389. Maiyalagan, T. Pt–Ru nanoparticles supported PAMAM dendrimer functionalized carbon nanofiber composite catalysts and their application to methanol oxidation. J. Solid State Electrochem. 2009, 13, 1561–1566. [Google Scholar] [CrossRef]
  390. Boronoev, M.P.; Zolotukhina, A.V.; Ignatyeva, V.I.; Terenina, M.V.; Maximov, A.L.; Karakhanov, E.A. Palladium catalysts based on mesoporous organic materials in semihydrogenation of alkynes. Macromol. Symp. 2016, 363, 57–63. [Google Scholar] [CrossRef]
  391. Meng, Y.; Gu, D.; Zhang, F.Q.; Shi, Y.F.; Yang, H.F.; Li, Z.; Yu, C.Z.; Tu, B.; Zhao, D.Y. Ordered mesoporous polymers and homologous carbon frameworks: Amphiphilic surfactant templating and direct transformation. Angew. Chem. Int. Ed. 2005, 44, 7053–7059. [Google Scholar] [CrossRef]
  392. Tonga, M.; Yesilbag Tonga, G.; Seber, G.; Gok, O.; Sanyal, A. Dendronized polystyrene via orthogonal double-click reactions. J. Polym. Sci. A Polym. Chem. 2013, 51, 5029–5037. [Google Scholar] [CrossRef]
  393. Karakhanov, E.A.; Maximov, A.L.; Zolotukhina, A.V.; Yatmanova, N.; Rosenberg, E. Alkyne hydrogenation using Pd–Ag hybrid nanocatalysts in surface-immobilized dendrimers. Appl. Organomet. Chem. 2015, 29, 777–784. [Google Scholar] [CrossRef]
  394. Karakhanov, E.A.; Boronoev, M.P.; Filippova, T.Y.; Maksimov, A.L. Guaiacol hydrogenation in an aqueous medium in the presence of a palladium catalyst supported on a mesoporous dendrimer-containing polymer. Petrol. Chem. 2018, 58, 407–411. [Google Scholar] [CrossRef]
  395. Hughes, M.A.; Nielsen, D.; Rosenberg, E.; Gobetto, R.; Viale, A.; Burton, S.D.; Ferel, J. Structural Investigations of silica polyamine composites: Surface coverage, metal ion coordination, and ligand modification. Ind. Eng. Chem. Res. 2006, 45, 6538–6547. [Google Scholar] [CrossRef]
  396. Karakhanov, E.; Maximov, A.; Zolotukhina, A.; Mamadli, A.; Vutolkina, A.; Ivanov, A. Dendrimer-stabilized Ru nanoparticles immobilized in organo-silica materials for hydrogenation of phenols. Catalysts 2017, 7, 86. [Google Scholar] [CrossRef]
  397. Garcia-Martinez, J.; Linares, N.; Sinibaldi, S.; Coronado, E.; Ribera, A. Incorporation of Pd nanoparticles in mesostructured silica. Microporous Mesoporous Mater. 2009, 117, 170–177. [Google Scholar] [CrossRef]
  398. Karaki, M.; Karout, A.; Toufaily, J.; Rataboul, F.; Essayem, N.; Lebeau, B. Synthesis and characterization of acidic ordered mesoporous organosilica SBA-15: Application to the hydrolysis of cellobiose and insight into the stability of the acidic functions. J. Catal. 2013, 305, 204–216. [Google Scholar] [CrossRef]
  399. Ishito, N.; Kobayashi, H.; Nakajima, K.; Maegawa, Y.; Inagaki, S.; Hara, K.; Fukuoka, A. Ruthenium-Immobilized Periodic Mesoporous Organosilica: Synthesis, Characterization, and Catalytic Application for Selective Oxidation of Alkanes. Chem. Eur. J. 2015, 21, 15564–15569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  400. Garcia, R.A.; van Grieken, R.; Iglesias, J.; Morales, V.; Villajos, N. Facile one-pot approach to the synthesis of chiral periodic mesoporous organosilicas SBA-15-type materials. J.Catal. 2010, 274, 221–227. [Google Scholar] [CrossRef]
  401. Maximov, A.L.; Zolotukhina, A.V.; Mamedli, A.A.; Kulikov, L.A.; Karakhanov, E.A. Selective levulinic acid hydrogenation in the presence of hybrid dendrimer-based catalysts. Part I: Monometallic. ChemCatChem 2018, 10, 222–233. [Google Scholar] [CrossRef]
  402. Nie, Y.; Jaenicke, S.; van Bekkum, H.; Chuah, G.-K. Stereoselective cascade hydrogenation of 4-tert-butylphenol and p-cresol over Zr-zeolite beta-supported rhodium. J. Catal. 2007, 246, 223–231. [Google Scholar] [CrossRef]
  403. Hiyoshi, N.; Rode, C.V.; Sato, O.; Tetsuka, H.; Shirai, M. Stereoselective hydrogenation of tert-butylphenols over charcoal-supported rhodium catalyst in supercritical carbon dioxide solvent. J. Catal. 2007, 252, 57–68. [Google Scholar] [CrossRef]
  404. Reetz, M.T.; Giebel, D. Cross-linked scandium-containing dendrimers: New class of heterogeneous catalysts. Angew. Chem. Int. Ed. 2000, 39, 2498–2501. [Google Scholar] [CrossRef]
  405. Badetti, E.; Caminade, A.-M.; Majoral, J.-P.; Moreno-Mañas, M.; Sebastián, R.M. Palladium (0) Nanoparticles stabilized by phosphorus dendrimers containing coordinating 15-membered triolefinic macrocycles in periphery. Langmuir 2008, 24, 2090–2101. [Google Scholar] [CrossRef]
  406. Schmid, G.; Emmrich, E.; Majoral, J.-P.; Caminade, A.-M. The behavior of Au55 nanoclusters on and in thiol-terminated dendrimer monolayers. Small 2005, 1, 73–75. [Google Scholar] [CrossRef]
  407. Alvarez, J.; Sun, L.; Crooks, R.M. Electroactive composite dendrimer films containing thiophene-terminated poly(amidoamine) dendrimers cross-linked by poly(3-methylthiophene). Chem. Mater. 2002, 14, 3995–4001. [Google Scholar] [CrossRef]
  408. Karakhanov, E.A.; Maximov, A.L.; Zolotukhina, A.V.; Terenina, M.V.; Vutolkina, A.V. Nanoheterogeneous ruthenium-containing catalysts based on dendrimers in the hydrogenation of aromatic compounds under two-phase conditions. Petrol. Chem. 2016, 56, 491–502. [Google Scholar] [CrossRef]
  409. Karakhanov, E.A.; Maksimov, A.L.; Zolotukhina, A.V.; Kardashev, S.V.; Filippova, T.Y. Palladium nanoparticles on dendrimer-containing supports as catalysts for hydrogenation of unsaturated hydrocarbons. Petrol. Chem. 2012, 52, 289–298. [Google Scholar] [CrossRef]
  410. Férey, G.; Serre, C. Large breathing effects in three-dimensional porous hybrid matter: Facts, analyses, rules and consequences. Chem. Soc. Rev. 2009, 38, 1380–1399. [Google Scholar] [CrossRef]
  411. Nalaparaju, A.; Jiang, J. Metal−organic frameworks for liquid phase applications. Adv. Sci. 2021, 8, 2003143. [Google Scholar] [CrossRef]
  412. Zakharyan, E.M.; Ma, G.; Maksimov, A.L.; Karakhanov, E.A.; Voronina, Z.D. Phenol and Dihydroxybenzene Hydrogenation catalysts based on polyamide dendrimers and rhodium species. Petrol. Chem. 2014, 54, 412–419. [Google Scholar] [CrossRef]
  413. Karakhanov, E.A.; Maksimov, A.L.; Zolotukhina, A.V.; Kardashev, S.V. Design of Dendrimer-based nanostructured catalyst systems and their catalytic activity in hydrogenation: Synthesis of Ruthenium nanoparticles immobilized in dendrimer networks. Petrol. Chem. 2010, 50, 290–297. [Google Scholar] [CrossRef]
  414. Sizova, I.A.; Kulikov, A.B.; Zolotukhina, A.V.; Serdyukov, S.I.; Maksimov, A.L.; Karakhanov, E.A. Synthesis of Ni–W aromatic hydrocarbon hydrogenation catalysts by the ex situ and in situ decomposition of a precursor based on a dendrimer network. Petrol. Chem. 2016, 56, 1107–1113. [Google Scholar] [CrossRef]
  415. Parenago, O.P.; Timashev, P.S.; Karakhanov, E.A.; Maximov, A.L.; Lazhko, A.E.; Zolotukhina, A.V.; Bagratashvili, V.N. Obtaining of highly-active catalysts of unsaturated compounds hydrogenation by using supercritical carbon dioxide. J. Supercrit. Fluids 2018, 140, 387–393. [Google Scholar] [CrossRef]
  416. Nabavinia, M.; Kanjilal, B.; Hesketh, A.; Wall, P.; Shirazi Amin, A.; Kerns, P.; Stanzione III, J.F.; Suib, S.; Liu, F.; Noshadi, I. One-pot aqueous and template-free synthesis of mesoporous polymeric resins. Catalysts 2019, 9, 782. [Google Scholar] [CrossRef] [Green Version]
  417. Ogasawara, S.; Kato, S. Palladium nanoparticles captured in microporous polymers: A tailor-made catalyst for heterogeneous carbon cross-coupling reactions. J. Am. Chem. Soc. 2010, 132, 4608–4613. [Google Scholar] [CrossRef]
  418. Shifrina, Z.B.; Bronstein, L.M. Magnetically recoverable catalysts: Beyond magnetic separation. Front. Chem. 2018, 6, 298. [Google Scholar] [CrossRef]
  419. Kainz, Q.M.; Reiser, O. Polymer- and dendrimer-coated magnetic nanoparticles as versatile supports for catalysts, scavengers, and reagents. Acc. Chem. Res. 2014, 47, 667–677. [Google Scholar] [CrossRef] [PubMed]
  420. Easterday, R.; Sanchez-Felix, O.; Losovyj, Y.; Pink, M.; Stein, B.D.; Morgan, D.G.; Rakitin, M.; Doluda, V.Y.; Sulman, M.G.; Mahmoud, W.E.; et al. Design of ruthenium/iron oxide nanoparticle mixtures for hydrogenation of nitrobenzene. Catal. Sci. Technol. 2015, 5, 1902–1910. [Google Scholar] [CrossRef]
  421. Oracko, T.; Jaquish, R.; Losovyj, Y.B.; Morgan, D.G.; Pink, M.; Stein, B.D.; Doluda, V.Y.; Tkachenko, O.P.; Shifrina, Z.B.; Grigoriev, M.E.; et al. Metal-ion distribution and oxygen vacancies that determine the activity of magnetically recoverable catalysts in methanol synthesis. ACS Appl. Mater. Interfaces 2017, 9, 34005–34014. [Google Scholar] [CrossRef] [PubMed]
  422. Manaenkov, O.V.; Mann, J.J.; Kislitza, O.V.; Losovyj, Y.; Stein, B.D.; Morgan, D.G.; Pink, M.; Lependina, O.L.; Shifrina, Z.B.; Matveeva, V.G.; et al. Ru-containing magnetically recoverable catalysts: A sustainable pathway from cellulose to ethylene and propylene glycols. ACS Appl. Mater. Interfaces 2016, 8, 21285–21293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  423. Mann, J.; Doluda, V.Y.; Leonard, C.; Losovyj, Y.B.; Morgan, D.G.; Bukalov, S.S.; Shifrina, Z.; Stein, B.D.; Cherkasov, N.; Rebrov, E.V.; et al. Metal oxide-zeolite composites in transformation of methanol to hydrocarbons: Do iron oxide and nickel oxide matter? RSC Adv. 2016, 6, 75166–75177. [Google Scholar] [CrossRef]
  424. Knecht, M.R.; Garcia-Martinez, J.C.; Crooks, R.M. Synthesis, Characterization, and magnetic properties of dendrimer-encapsulated nickel nanoparticles containing <150 Atoms. Chem. Mater. 2006, 18, 5039–5044. [Google Scholar] [CrossRef]
  425. Duanmu, C.; Saha, I.; Zheng, Y.; Goodson, B.M.; Gao, Y. Dendron-functionalized superparamagnetic nanoparticles with switchable solubility in organic and aqueous media:  Matrices for homogeneous catalysis and potential MRI contrast agents. Chem. Mater. 2006, 18, 5973–5981. [Google Scholar] [CrossRef]
  426. Rehm, T.H.; Bogdan, A.; Hofmann, C.; Löb, P.; Shifrina, Z.B.; Morgan, D.G.; Bronstein, L.M. Proof of concept: Magnetic fixation of dendron-functionalized iron oxide nanoparticles containing palladium nanoparticles for continuous-flow Suzuki Coupling reactions. ACS Appl. Mater. Interfaces 2015, 7, 27254–27261. [Google Scholar] [CrossRef] [Green Version]
  427. Kuchkina, N.V.; Haskell, A.K.; Sorokina, S.A.; Torozova, A.S.; Nikoshvili, L.Z.; Sulman, E.M.; Stein, B.D.; Gene Morgan, D.; Bronstein, L.M.; Shifrina, Z.B. Pd Catalyst based on hyperbranched polypyridylphenylene formed in situ on magnetic silica allows for excellent performance in Suzuki−Miyaura reaction. ACS Appl. Mater. Interfaces 2020, 12, 22170–22178. [Google Scholar] [CrossRef]
  428. Yuzik-Klimova, E.Y.; Kuchkina, N.V.; Sorokina, S.A.; Morgan, D.G.; Boris, B.; Nikoshvili, L.Z.; Lyubimova, N.A.; Matveeva, V.G.; Sulman, E.M.; Stein, B.D.; et al. Magnetically recoverable catalysts based on polyphenylenepyridyl dendrons and dendrimers. RSC Adv. 2014, 4, 23271–23280. [Google Scholar] [CrossRef]
  429. Kuchkina, N.V.; Yuzik-Klimova, E.Y.u.; Sorokina, S.A.; Peregudov, A.S.; Antonov, D.Y.; Gage, S.H.; Boris, B.S.; Nikoshvili, L.Z.; Sulman, E.M.; Morgan, D.G.; et al. Polyphenylenepyridyl Dendrons with Functional Periphery and Focal Points: Syntheses and Applications. Macromolecules 2013, 46, 5890–5898. [Google Scholar] [CrossRef]
  430. Rosario-Amorin, D.; Gaboyard, M.; Clérac, R.; Vellutini, L.; Nlate, S.; Heuzé, K. Metallodendritic Grafted Core–Shell γ-Fe2O3 Nanoparticles used as recoverable catalysts in Suzuki C–C coupling reactions. Chem. Eur. J. 2012, 18, 3305–3315. [Google Scholar] [CrossRef]
  431. Wang, D.; Deraedt, C.; Ruiz, J.; Astruc, D. Magnetic and dendritic catalysts. Acc. Chem. Res. 2015, 48, 7–1871. [Google Scholar] [CrossRef]
  432. Wang, D.; Deraedt, C.; Salmon, L.; Labrugère, C.; Etienne, L.; Ruiz, J.; Astruc, D. Efficient and Magnetically recoverable “click” PEGylated γ-Fe2O3–Pd nanoparticle catalysts for Suzuki–Miyaura, Sonogashira, and Heck Reactions with positive dendritic effects. Chem. Eur. J. 2015, 21, 1508–1519. [Google Scholar] [CrossRef]
  433. Deraedt, C.; Wang, D.; Salmon, L.; Etienne, L.; Labrugère, C.; Ruiz, J.; Astruc, D. Robust, efficient and recyclable catalyst by impregnation of dendritically preformed Pd nanoparticles on magnetic support. ChemCatChem 2015, 7, 303–308. [Google Scholar] [CrossRef]
  434. Abu-Reziq, R.; Alper, H.; Wang, D.; Post, M.L. Metal supported on dendronized magnetic nanoparticles: Highly selective hydroformylation catalysts. J. Am. Chem. Soc. 2006, 128, 5279–5282. [Google Scholar] [CrossRef] [Green Version]
  435. Fujita, K.-I.; Umeki, S.; Yamazaki, M.; Ainoya, T.; Tsuchimoto, T.; Yasuda, H. Magnetically recoverable osmium catalysts for dihydroxylation of olefins. Tetrahedron Lett. 2011, 52, 3137–3140. [Google Scholar] [CrossRef]
  436. Keller, M.; Collière, V.; Reiser, O.; Caminade, A.-M.; Majoral, J.-P.; Ouali, A. Pyrene-tagged dendritic catalysts noncovalently grafted onto magnetic Co/C nanoparticles: An efficient and recyclable system for drug synthesis. Angew. Chem. Int. Ed. 2013, 52, 3626–3629. [Google Scholar] [CrossRef]
Figure 1. Structures of some dendrimer types, used in nanocatalysis [10,63,64,71,72,90,93].
Figure 1. Structures of some dendrimer types, used in nanocatalysis [10,63,64,71,72,90,93].
Polymers 14 00981 g001
Figure 2. Dendrimer structure on the example of PPI dendrimer [26,71]. Adapted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Figure 2. Dendrimer structure on the example of PPI dendrimer [26,71]. Adapted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Polymers 14 00981 g002
Figure 3. Possible locations of active metallocenters in dendrimer: (a) core; (b) nodes; (c) end groups; (d) inner cavity [57,61,78,90,91,92,135].
Figure 3. Possible locations of active metallocenters in dendrimer: (a) core; (b) nodes; (c) end groups; (d) inner cavity [57,61,78,90,91,92,135].
Polymers 14 00981 g003
Scheme 1. The typical approach for two-step synthesis of dendrimer-encapsulated nanoparticles (DENs), using PAMAM, PPI, and other poly(alkyl amine) dendrimers as nanoreactor stabilizing ligands [57,91,115,116,117,151,152,153,154,155,156,157].
Scheme 1. The typical approach for two-step synthesis of dendrimer-encapsulated nanoparticles (DENs), using PAMAM, PPI, and other poly(alkyl amine) dendrimers as nanoreactor stabilizing ligands [57,91,115,116,117,151,152,153,154,155,156,157].
Polymers 14 00981 sch001
Scheme 2. Possible pathways for the synthesis of various type bimetallic DENs [88,95,115].
Scheme 2. Possible pathways for the synthesis of various type bimetallic DENs [88,95,115].
Polymers 14 00981 sch002
Scheme 3. Radial complexation of poly(phenylazomethine) dendrimer with different metal ions and their further reduction to alloy-type nanoparticle [10,80,166,183,184,185].
Scheme 3. Radial complexation of poly(phenylazomethine) dendrimer with different metal ions and their further reduction to alloy-type nanoparticle [10,80,166,183,184,185].
Polymers 14 00981 sch003
Figure 4. Dendrimer-stabilized (left) and dendrimer-encapsulated nanoparticles (right) [25,90,150,186,187,188,191,193].
Figure 4. Dendrimer-stabilized (left) and dendrimer-encapsulated nanoparticles (right) [25,90,150,186,187,188,191,193].
Polymers 14 00981 g004
Figure 5. Pd-cored nanoparticle, coated with thiol poly(aryl ether) dendrons [26,121]. Reprinted (adapted) with permission of American Chemical Society, from: Gopidas, K. R.; Whitesell, J. K.; Fox, M. A. Synthesis, Characterization, and Catalytic Applications of a Palladium-Nanoparticle-Cored Dendrimer. Nano Lett. 2003, 3, 1757–1760 [121]. Copyright 2003, American Chemical Society.
Figure 5. Pd-cored nanoparticle, coated with thiol poly(aryl ether) dendrons [26,121]. Reprinted (adapted) with permission of American Chemical Society, from: Gopidas, K. R.; Whitesell, J. K.; Fox, M. A. Synthesis, Characterization, and Catalytic Applications of a Palladium-Nanoparticle-Cored Dendrimer. Nano Lett. 2003, 3, 1757–1760 [121]. Copyright 2003, American Chemical Society.
Polymers 14 00981 g005
Scheme 4. NCD synthesis via post-functionalization method [62,201]. Reprinted (adapted) with permission of American Chemical Society, from: Shon, Y.-S.; Choi, D.; Dare, J.; Dinh, T. Synthesis of Nanoparticle-Cored Dendrimers by Convergent Dendritic Functionalization of Monolayer-Protected Nanoparticles. Langmuir 2008, 24, 6924–6931 [201]. Copyright 2008, American Chemical Society.
Scheme 4. NCD synthesis via post-functionalization method [62,201]. Reprinted (adapted) with permission of American Chemical Society, from: Shon, Y.-S.; Choi, D.; Dare, J.; Dinh, T. Synthesis of Nanoparticle-Cored Dendrimers by Convergent Dendritic Functionalization of Monolayer-Protected Nanoparticles. Langmuir 2008, 24, 6924–6931 [201]. Copyright 2008, American Chemical Society.
Polymers 14 00981 sch004
Figure 6. Pt-cored nanoparticle, coated with tris(acetic acid) ammonium poly(aryl ether) dendrons [62,200]. Reprinted with permission of Elsevier, from: Yang, P.; Zhang, W.; Du, Y.; Wang, X. Hydrogenation of nitrobenzenes catalyzed by platinum nanoparticle core-polyaryl ether trisacetic acid ammonium chloride dendrimer shell nanocomposite. J. Mol. Catal. A: Chem. 2006, 260, 4–10 [200]. Copyright 2006, Elsevier.
Figure 6. Pt-cored nanoparticle, coated with tris(acetic acid) ammonium poly(aryl ether) dendrons [62,200]. Reprinted with permission of Elsevier, from: Yang, P.; Zhang, W.; Du, Y.; Wang, X. Hydrogenation of nitrobenzenes catalyzed by platinum nanoparticle core-polyaryl ether trisacetic acid ammonium chloride dendrimer shell nanocomposite. J. Mol. Catal. A: Chem. 2006, 260, 4–10 [200]. Copyright 2006, Elsevier.
Polymers 14 00981 g006
Figure 7. Au nanoparticle, stabilized by thiol poly(L-lysine) dendrons [84,204].
Figure 7. Au nanoparticle, stabilized by thiol poly(L-lysine) dendrons [84,204].
Polymers 14 00981 g007
Figure 8. CdSe/ZnS nanoparticle, stabilized by thiol PAMAM dendrons [84]. Reprinted and adapted with permission of Elsevier, from: Smith, D. K.; Hirst, A. R.; Love, C. S.; Hardy, J. G.; Brignell, S. V.; Huang, B. Self-assembly using dendritic building blocks—towards controllable nanomaterials. Prog. Polym. Sci. 2005, 30, 220–293 [84]. Copyright 2005, Elsevier.
Figure 8. CdSe/ZnS nanoparticle, stabilized by thiol PAMAM dendrons [84]. Reprinted and adapted with permission of Elsevier, from: Smith, D. K.; Hirst, A. R.; Love, C. S.; Hardy, J. G.; Brignell, S. V.; Huang, B. Self-assembly using dendritic building blocks—towards controllable nanomaterials. Prog. Polym. Sci. 2005, 30, 220–293 [84]. Copyright 2005, Elsevier.
Polymers 14 00981 g008
Scheme 5. The synthesis of bimetallic PtPd nanoparticles, stabilized by PAMAM dendrons with 4-(aminomethyl pyridine) in focal point [131,207]. Adapted with permission of The Chemical Society of Japan, from: Murata, M.; Tanaka, Yu.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Palladium–Platinum Bimetallic Nanoparticle Catalysts Using Dendron Assembly for Selective Hydrogenation of Dienes and Their Application to Thermomorphic System. Chem. Lett. 2005, 34, 272–273 [207]. Copyright The Chemical Society of Japan, 2005.
Scheme 5. The synthesis of bimetallic PtPd nanoparticles, stabilized by PAMAM dendrons with 4-(aminomethyl pyridine) in focal point [131,207]. Adapted with permission of The Chemical Society of Japan, from: Murata, M.; Tanaka, Yu.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Palladium–Platinum Bimetallic Nanoparticle Catalysts Using Dendron Assembly for Selective Hydrogenation of Dienes and Their Application to Thermomorphic System. Chem. Lett. 2005, 34, 272–273 [207]. Copyright The Chemical Society of Japan, 2005.
Polymers 14 00981 sch005
Scheme 6. Immobilization of dendrimer-encapsulated nanoparticles inside a mesoporous carrier [292]. Reprinted and adapted with permission of American Chemical Society, from: Huang, W.; Kuhn, J. N.; Tsung, Ch.-K.; Zhang, Ya.; Habas, S. E.; Yang, P.; Somorjai, G. A. Dendrimer Templated Synthesis of One Nanometer Rh and Pt Particles Supported on Mesoporous Silica: Catalytic Activity for Ethylene and Pyrrole Hydrogenation. Nano Lett. 2008, 8, 2027–2034 [290]. Copyright 2008, American Chemical Society.
Scheme 6. Immobilization of dendrimer-encapsulated nanoparticles inside a mesoporous carrier [292]. Reprinted and adapted with permission of American Chemical Society, from: Huang, W.; Kuhn, J. N.; Tsung, Ch.-K.; Zhang, Ya.; Habas, S. E.; Yang, P.; Somorjai, G. A. Dendrimer Templated Synthesis of One Nanometer Rh and Pt Particles Supported on Mesoporous Silica: Catalytic Activity for Ethylene and Pyrrole Hydrogenation. Nano Lett. 2008, 8, 2027–2034 [290]. Copyright 2008, American Chemical Society.
Polymers 14 00981 sch006
Figure 9. Reactions, catalyzed by PAMAM dendrimer-encapsulated nanoparticles, immobilized in the channels of SBA-15 mesoporous material [64,293].
Figure 9. Reactions, catalyzed by PAMAM dendrimer-encapsulated nanoparticles, immobilized in the channels of SBA-15 mesoporous material [64,293].
Polymers 14 00981 g009
Figure 10. A supposed intermediate structure in Hayashi-Ito aldol cycloaddition to isocyanoacetate in the presence of PAMAM dendrimer-encapsulated gold nanoparticles, immobilized in the channels of SBA-15 mesoporous material [311].
Figure 10. A supposed intermediate structure in Hayashi-Ito aldol cycloaddition to isocyanoacetate in the presence of PAMAM dendrimer-encapsulated gold nanoparticles, immobilized in the channels of SBA-15 mesoporous material [311].
Polymers 14 00981 g010
Scheme 7. Hydrogenation of 2-nitrobenzaldehyde in the presence of Ir@PAMAM-G6/SiO2 [227].
Scheme 7. Hydrogenation of 2-nitrobenzaldehyde in the presence of Ir@PAMAM-G6/SiO2 [227].
Polymers 14 00981 sch007
Scheme 8. Possible pathways of citral hydrogenation [312].
Scheme 8. Possible pathways of citral hydrogenation [312].
Polymers 14 00981 sch008
Figure 11. Selectivity vs. conversion plots for Ru@PAMAM/SiO2 SCILL catalysts in the presence of [EMIM][NTf2] ionic liquid at 110 °C, 30 atm. of H2 [313]: (a) in the beginning of the reaction (0–10 min.); (b) at conversions of 9–13%, 37–52%, and 69–73%.
Figure 11. Selectivity vs. conversion plots for Ru@PAMAM/SiO2 SCILL catalysts in the presence of [EMIM][NTf2] ionic liquid at 110 °C, 30 atm. of H2 [313]: (a) in the beginning of the reaction (0–10 min.); (b) at conversions of 9–13%, 37–52%, and 69–73%.
Polymers 14 00981 g011
Figure 12. Reactions, catalyzed by DPA-encapsulated subnanoclusters, immobilized on zirconia or carbon supports [10,64].
Figure 12. Reactions, catalyzed by DPA-encapsulated subnanoclusters, immobilized on zirconia or carbon supports [10,64].
Polymers 14 00981 g012
Figure 13. Immobilization of PAMAM dendrimer-encapsulated palladium nanoparticles on the gold surface via electrostatic interactions, using a non-covalent MUA linker [26,92]. Reprinted by permission of Springer Nature, from: Karakhanov, E.A.; Maksimov, A.L.; Zolotukhina, A.V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Figure 13. Immobilization of PAMAM dendrimer-encapsulated palladium nanoparticles on the gold surface via electrostatic interactions, using a non-covalent MUA linker [26,92]. Reprinted by permission of Springer Nature, from: Karakhanov, E.A.; Maksimov, A.L.; Zolotukhina, A.V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Polymers 14 00981 g013
Scheme 9. Adsorption of PAMAM dendrimer-encapsulated Pt NPs on the nitrogen-doped carbon nanotubes [329].
Scheme 9. Adsorption of PAMAM dendrimer-encapsulated Pt NPs on the nitrogen-doped carbon nanotubes [329].
Polymers 14 00981 sch009
Scheme 10. PAMAM dendrimer-encapsulated Au nanoparticles, immobilized on the gold support via linker: hexa(ethylene glycol) carboxylic acid-terminated thiol, modified N-(ε-maleimido-caproic acid) hydrazide [62,330]. Reprinted and adapted with permission of American Chemical Society, from: Frasconi, M.; Tortolini, C.; Botre, F.; Mazzei, F. Multifunctional Au Nanoparticle Dendrimer-Based Surface Plasmon Resonance Biosensor and Its Application for Improved Insulin Detection. Anal. Chem. 2010, 82, 7335–7342 [330]. Copyright 2010, American Chemical Society.
Scheme 10. PAMAM dendrimer-encapsulated Au nanoparticles, immobilized on the gold support via linker: hexa(ethylene glycol) carboxylic acid-terminated thiol, modified N-(ε-maleimido-caproic acid) hydrazide [62,330]. Reprinted and adapted with permission of American Chemical Society, from: Frasconi, M.; Tortolini, C.; Botre, F.; Mazzei, F. Multifunctional Au Nanoparticle Dendrimer-Based Surface Plasmon Resonance Biosensor and Its Application for Improved Insulin Detection. Anal. Chem. 2010, 82, 7335–7342 [330]. Copyright 2010, American Chemical Society.
Polymers 14 00981 sch010
Figure 14. PAMAM dendron-assisted non-covalent immobilization of Mn2+/Co2+ catalyst for p-xylene oxidation on the bentonite surface [324,331]. Reprinted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Vinokurov, V. A. Oxidation of p-Xylene. A Review. Russ. J. Appl. Chem. 2018, 91, 707−727 [324]. Copyright 2018, Springer Nature.
Figure 14. PAMAM dendron-assisted non-covalent immobilization of Mn2+/Co2+ catalyst for p-xylene oxidation on the bentonite surface [324,331]. Reprinted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Vinokurov, V. A. Oxidation of p-Xylene. A Review. Russ. J. Appl. Chem. 2018, 91, 707−727 [324]. Copyright 2018, Springer Nature.
Polymers 14 00981 g014
Scheme 11. Covalent dendron immobilization on silica surface via divergent approach on the example of PAMAM dendrimer [26,86,92,148,332]. Reprinted by permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Scheme 11. Covalent dendron immobilization on silica surface via divergent approach on the example of PAMAM dendrimer [26,86,92,148,332]. Reprinted by permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Polymers 14 00981 sch011
Scheme 12. The synthesis of Mn (II) salen complex catalyst based on silica-grafted PAMAM dendron [252].
Scheme 12. The synthesis of Mn (II) salen complex catalyst based on silica-grafted PAMAM dendron [252].
Polymers 14 00981 sch012
Figure 15. G3.5 (1R, 2S)-ephedrine-terminated PAMAM dendron, covalently immobilized on the silica surface [92]. Adapted with permission of Springer, from: Chung, You.-M.; Rhee, H.-K. Internal/External Use of Dendrimer in Catalysis. Korean J. Chem. Eng. 2004, 21, 81–97 [92]. Copyright Springer, 2004.
Figure 15. G3.5 (1R, 2S)-ephedrine-terminated PAMAM dendron, covalently immobilized on the silica surface [92]. Adapted with permission of Springer, from: Chung, You.-M.; Rhee, H.-K. Internal/External Use of Dendrimer in Catalysis. Korean J. Chem. Eng. 2004, 21, 81–97 [92]. Copyright Springer, 2004.
Polymers 14 00981 g015
Scheme 13. The synthesis of dendrimer-based hollow capsule material [26,334]. Adapted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Scheme 13. The synthesis of dendrimer-based hollow capsule material [26,334]. Adapted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Polymers 14 00981 sch013
Figure 16. Heterogeneous dendritic catalysts, based on phosphine (left) and PCP complex-terminated PAMAM dendrons (right), covalently immobilized on the silica surface, suggested by Alper et al. [93,149,334]. Adapted with permission of John Wiley and Sons, from: Chanthateyanonth, R.; Alper, H. Recyclable Tridentate Stable Palladium (II) PCP-Type Catalysts Supported on Silica for the Selective Synthesis of Lactones. Adv. Synth. Catal. 2004, 346, 1375–1384 [334]. Copyright John Wiley and Sons, 2004.
Figure 16. Heterogeneous dendritic catalysts, based on phosphine (left) and PCP complex-terminated PAMAM dendrons (right), covalently immobilized on the silica surface, suggested by Alper et al. [93,149,334]. Adapted with permission of John Wiley and Sons, from: Chanthateyanonth, R.; Alper, H. Recyclable Tridentate Stable Palladium (II) PCP-Type Catalysts Supported on Silica for the Selective Synthesis of Lactones. Adv. Synth. Catal. 2004, 346, 1375–1384 [334]. Copyright John Wiley and Sons, 2004.
Polymers 14 00981 g016
Scheme 14. Product distributions in the hydrogenation of dienes at full substrate conversions in the presence of silica-supported catalysts based on grafted PAMAM dendrons terminated with Pd phosphine complexes. Reaction conditions: 25 °C, 1 atm. of H2, MeOH, 0.5–24 h [148].
Scheme 14. Product distributions in the hydrogenation of dienes at full substrate conversions in the presence of silica-supported catalysts based on grafted PAMAM dendrons terminated with Pd phosphine complexes. Reaction conditions: 25 °C, 1 atm. of H2, MeOH, 0.5–24 h [148].
Polymers 14 00981 sch014
Scheme 15. Carbonylation of: (a) otho-allylphenols in the presence of SiO2-PAMAM-PCP-Pd complex catalyst [93,334]; (b) otho-vinylphenols in the presence of SiO2-PAMAM-PPh2-Pd complex catalyst [93,336]; (c) 2-(2-iodoaryloxy)anilines and other related substrates in the presence of SiO2-PAMAM-PPh2-Pd complex catalyst [93,337].
Scheme 15. Carbonylation of: (a) otho-allylphenols in the presence of SiO2-PAMAM-PCP-Pd complex catalyst [93,334]; (b) otho-vinylphenols in the presence of SiO2-PAMAM-PPh2-Pd complex catalyst [93,336]; (c) 2-(2-iodoaryloxy)anilines and other related substrates in the presence of SiO2-PAMAM-PPh2-Pd complex catalyst [93,337].
Polymers 14 00981 sch015
Figure 17. The structure of nano-SiO2-PAMAM-Pd/IL catalyst (left) [26], and its raspberry-shaped ensembles in TEM image (right) [343]. Reprinted by permissions of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer; and of Georg Thieme Verlag KG, from Hagiwara, H.; Sasaki, H.; Tsubokawa, N.; Hoshi, T.; Suzuki, T.; Tsuda, T.; Kuwabata, S. Immobilization of Pd on Nanosilica Dendrimer as SILC: Highly Active and Sustainable Cluster Catalyst for Suzuki-Miyaura Reaction. Synlett 2010, 1990–1996 [343]. Copyright 2010, Georg Thieme Verlag KG.
Figure 17. The structure of nano-SiO2-PAMAM-Pd/IL catalyst (left) [26], and its raspberry-shaped ensembles in TEM image (right) [343]. Reprinted by permissions of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer; and of Georg Thieme Verlag KG, from Hagiwara, H.; Sasaki, H.; Tsubokawa, N.; Hoshi, T.; Suzuki, T.; Tsuda, T.; Kuwabata, S. Immobilization of Pd on Nanosilica Dendrimer as SILC: Highly Active and Sustainable Cluster Catalyst for Suzuki-Miyaura Reaction. Synlett 2010, 1990–1996 [343]. Copyright 2010, Georg Thieme Verlag KG.
Polymers 14 00981 g017
Scheme 16. The synthesis DABCO-terminated PEI dendron, grafted to silica, and subsequent Pd nanoparticle impregnation [341].
Scheme 16. The synthesis DABCO-terminated PEI dendron, grafted to silica, and subsequent Pd nanoparticle impregnation [341].
Polymers 14 00981 sch016
Scheme 17. Cross-coupling reactions in the presence of SiO2-PEI-DABCO-Pd catalyst [341]: (a) Suzuki; (b) Heck.
Scheme 17. Cross-coupling reactions in the presence of SiO2-PEI-DABCO-Pd catalyst [341]: (a) Suzuki; (b) Heck.
Polymers 14 00981 sch017
Scheme 18. The synthesis triazine—propylene imine dendron, grafted to nano-silica (nSTDP), and subsequent Pd nanoparticle impregnation [342,345].
Scheme 18. The synthesis triazine—propylene imine dendron, grafted to nano-silica (nSTDP), and subsequent Pd nanoparticle impregnation [342,345].
Polymers 14 00981 sch018
Scheme 19. C–C cross-coupling reactions in the presence of Pd-nSTDP catalyst [342]: (a) Suzuki; (b) Heck.
Scheme 19. C–C cross-coupling reactions in the presence of Pd-nSTDP catalyst [342]: (a) Suzuki; (b) Heck.
Polymers 14 00981 sch019
Scheme 20. C–S cross-coupling reactions in the presence of Pd-nSTDP catalyst [345].
Scheme 20. C–S cross-coupling reactions in the presence of Pd-nSTDP catalyst [345].
Polymers 14 00981 sch020
Scheme 21. An assumed mechanism for C–S cross-coupling of aryl halides and disulfides, providing the effective reaction proceeding [345].
Scheme 21. An assumed mechanism for C–S cross-coupling of aryl halides and disulfides, providing the effective reaction proceeding [345].
Polymers 14 00981 sch021
Scheme 22. Grafting of PAMAM dendrons inside the channels of ordered mesoporous materials SBA-15 or MCM-41 [26,352,353]. Reprinted and adapted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Scheme 22. Grafting of PAMAM dendrons inside the channels of ordered mesoporous materials SBA-15 or MCM-41 [26,352,353]. Reprinted and adapted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Polymers 14 00981 sch022
Scheme 23. Grafting of melamine-aminomethylpiperidine dendrons inside the channels of SBA-15 material and subsequent metal nanoparticle encapsulation [233].
Scheme 23. Grafting of melamine-aminomethylpiperidine dendrons inside the channels of SBA-15 material and subsequent metal nanoparticle encapsulation [233].
Polymers 14 00981 sch023
Scheme 24. An assumed mechanism for Cu-catalyzed heterogeneous alkyne-azide click cycloaddition [233].
Scheme 24. An assumed mechanism for Cu-catalyzed heterogeneous alkyne-azide click cycloaddition [233].
Polymers 14 00981 sch024
Scheme 25. Polystyrene activation and subsequent PAMAM dendron grafting [362,363,364,365,366,367].
Scheme 25. Polystyrene activation and subsequent PAMAM dendron grafting [362,363,364,365,366,367].
Polymers 14 00981 sch025
Figure 18. HRTEM images of the PS-PAMAM-G3 supported palladium nanoparticles [364]. Reprinted with permission of Taylor and Francis Academic Journals, from: Krishnan, G.R.; Sreekumar, K. Characterization of Polystyrene Supported Catalytically Active Poly(amidoamine) Dendrimer-Palladium Nanoparticle Conjugates. Soft Mater. 2010, 8, 114–129 [364]. Copyright 2010, Taylor & Francis Academic Journals.
Figure 18. HRTEM images of the PS-PAMAM-G3 supported palladium nanoparticles [364]. Reprinted with permission of Taylor and Francis Academic Journals, from: Krishnan, G.R.; Sreekumar, K. Characterization of Polystyrene Supported Catalytically Active Poly(amidoamine) Dendrimer-Palladium Nanoparticle Conjugates. Soft Mater. 2010, 8, 114–129 [364]. Copyright 2010, Taylor & Francis Academic Journals.
Polymers 14 00981 g018
Scheme 26. The synthesis of Au NPs, immobilized on polystyrene beads, coated with PAMAM dendrons [367].
Scheme 26. The synthesis of Au NPs, immobilized on polystyrene beads, coated with PAMAM dendrons [367].
Polymers 14 00981 sch026
Scheme 27. Benzyl alcohol oxidation and arylboronic acid homocoupling in the presence of PS-PAMAM-G3-Au catalyst [367].
Scheme 27. Benzyl alcohol oxidation and arylboronic acid homocoupling in the presence of PS-PAMAM-G3-Au catalyst [367].
Polymers 14 00981 sch027
Scheme 28. The synthesis of polystyrene-supported poly(aryl amide) dendrons and their subsequent modification with Rh phosphine complexes [369].
Scheme 28. The synthesis of polystyrene-supported poly(aryl amide) dendrons and their subsequent modification with Rh phosphine complexes [369].
Polymers 14 00981 sch028
Scheme 29. Grafting of Fréchet type poly(aryl ether) dendrons on Wang resin and their subsequent modification with Co or Pd phosphine complexes [370,371,373,374]: (a) via Mitsunobu reaction and successive ester reduction [372]; (b) via Mitsunobu reaction and after ester hydrolysis [373].
Scheme 29. Grafting of Fréchet type poly(aryl ether) dendrons on Wang resin and their subsequent modification with Co or Pd phosphine complexes [370,371,373,374]: (a) via Mitsunobu reaction and successive ester reduction [372]; (b) via Mitsunobu reaction and after ester hydrolysis [373].
Polymers 14 00981 sch029
Scheme 30. Pauson–Khand [2+2+1] cycloaddition in the presence Co poly(aryl ether) dendritic catalysts [374].
Scheme 30. Pauson–Khand [2+2+1] cycloaddition in the presence Co poly(aryl ether) dendritic catalysts [374].
Polymers 14 00981 sch030
Scheme 31. Heck cross-coupling in the presence Pd poly(aryl ether) dendritic catalysts [370].
Scheme 31. Heck cross-coupling in the presence Pd poly(aryl ether) dendritic catalysts [370].
Polymers 14 00981 sch031
Figure 19. Three-dimensional structure of polyaromatic frameworks [375,376,377,378,379]. Reprinted with permission of Elsevier, from: Karakhanov, E.; Maximov, A.; Terenina, M.; Vinokurov, V.; Kulikov, L.; Makeeva, D.; Glotov, A. Selective hydrogenation of terminal alkynes over palladium nanoparticles within the pores of amino-modified porous aromatic frameworks. Catal. Today 2020, 357, 176–184 [379]. Copyright Elsevier, 2020.
Figure 19. Three-dimensional structure of polyaromatic frameworks [375,376,377,378,379]. Reprinted with permission of Elsevier, from: Karakhanov, E.; Maximov, A.; Terenina, M.; Vinokurov, V.; Kulikov, L.; Makeeva, D.; Glotov, A. Selective hydrogenation of terminal alkynes over palladium nanoparticles within the pores of amino-modified porous aromatic frameworks. Catal. Today 2020, 357, 176–184 [379]. Copyright Elsevier, 2020.
Polymers 14 00981 g019
Scheme 32. The grafting of OH-terminated PEI dendrons on the PAF surface [363].
Scheme 32. The grafting of OH-terminated PEI dendrons on the PAF surface [363].
Polymers 14 00981 sch032
Figure 20. Hydrogenation of unsaturated compounds in the presence of Pd-PAF-PEI catalysts. Reaction conditions are: 80 °C, H2 10 atm., 30 min. [363].
Figure 20. Hydrogenation of unsaturated compounds in the presence of Pd-PAF-PEI catalysts. Reaction conditions are: 80 °C, H2 10 atm., 30 min. [363].
Polymers 14 00981 g020
Scheme 33. The synthesis of Pd nanoparticles, encapsulated on PAMAM dendrons, grafted to multiwall carbon nanotubes [26,372]. Adapted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Scheme 33. The synthesis of Pd nanoparticles, encapsulated on PAMAM dendrons, grafted to multiwall carbon nanotubes [26,372]. Adapted with permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Polymers 14 00981 sch033
Scheme 34. The synthesis of PAMAM dendron-functionalized single wall carbon nanotubes via convergent strategy and subsequent Pd nanoparticles encapsulation [380,382].
Scheme 34. The synthesis of PAMAM dendron-functionalized single wall carbon nanotubes via convergent strategy and subsequent Pd nanoparticles encapsulation [380,382].
Polymers 14 00981 sch034
Scheme 35. Cross-coupling reactions in the presence of SWCNT-PAMAM-Pd catalysts [382]: (a) Suzuki; (b) Heck.
Scheme 35. Cross-coupling reactions in the presence of SWCNT-PAMAM-Pd catalysts [382]: (a) Suzuki; (b) Heck.
Polymers 14 00981 sch035
Scheme 36. Conception of “release and catch” mechanism on the example Suzuki cross-coupling reaction, catalyzed by PAMAM dendrimer-encapsulated nanoparticles, immobilized on SWNT [10,64,382]. Adapted with permission of American Chemical Society, from: Giacalone, F.; Campisciano, V.; Calabrese, C.; La Parola, V.; Syrgiannis, Z.; Prato, M.; Gruttadauria, M. Single-Walled Carbon Nanotube–Polyamidoamine Dendrimer Hybrids for Heterogeneous Catalysis. ACS Nano 2016, 10, 4627–4636 [382]. Copyright American Chemical Society, 2016.
Scheme 36. Conception of “release and catch” mechanism on the example Suzuki cross-coupling reaction, catalyzed by PAMAM dendrimer-encapsulated nanoparticles, immobilized on SWNT [10,64,382]. Adapted with permission of American Chemical Society, from: Giacalone, F.; Campisciano, V.; Calabrese, C.; La Parola, V.; Syrgiannis, Z.; Prato, M.; Gruttadauria, M. Single-Walled Carbon Nanotube–Polyamidoamine Dendrimer Hybrids for Heterogeneous Catalysis. ACS Nano 2016, 10, 4627–4636 [382]. Copyright American Chemical Society, 2016.
Polymers 14 00981 sch036
Scheme 37. Surface modification of multiwalled carbon nanotubes with poly(oxymethylpyrogallol) dendrons [86,385].
Scheme 37. Surface modification of multiwalled carbon nanotubes with poly(oxymethylpyrogallol) dendrons [86,385].
Polymers 14 00981 sch037
Scheme 38. The anchoring of poly(ester amide) dendron of the 3rd generation to single-wall carbon nanotube and subsequent CdS nanoparticle impregnation [25,62,386].
Scheme 38. The anchoring of poly(ester amide) dendron of the 3rd generation to single-wall carbon nanotube and subsequent CdS nanoparticle impregnation [25,62,386].
Polymers 14 00981 sch038
Scheme 39. The synthesis of PAMAM dendrimer-encapsulated PtRu nanoparticles, immobilized on carbon nanofiber [389]. Adapted with permission of Springer Nature, from: Maiyalagan, T. Pt–Ru nanoparticles supported PAMAM dendrimer functionalized carbon nanofiber composite catalysts and their application to methanol oxidation. J. Solid State Electrochem. 2009, 13, 1561–1566 [389]. Copyright Springer, 2008.
Scheme 39. The synthesis of PAMAM dendrimer-encapsulated PtRu nanoparticles, immobilized on carbon nanofiber [389]. Adapted with permission of Springer Nature, from: Maiyalagan, T. Pt–Ru nanoparticles supported PAMAM dendrimer functionalized carbon nanofiber composite catalysts and their application to methanol oxidation. J. Solid State Electrochem. 2009, 13, 1561–1566 [389]. Copyright Springer, 2008.
Polymers 14 00981 sch039
Scheme 40. Immobilization of PPI dendrimers of the 3rd generation on the MPF surface [362].
Scheme 40. Immobilization of PPI dendrimers of the 3rd generation on the MPF surface [362].
Polymers 14 00981 sch040
Figure 21. TEM image (left) and particle size distribution (right) of the MPF-PPI-G3-Pd catalyst [362,390]. Reprinted with permission of John Wiley and Sons, from: Boronoev, M. P.; Zolotukhina, A. V.; Ignatyeva, V. I.; Terenina, M. V.; Maximov, A. L.; Karakhanov, E. A. Palladium Catalysts Based on Mesoporous Organic Materials in Semihydrogenation of Alkynes. Macromol. Symp. 2016, 363, 57–63 [390]. Copyright 2016, John Wiley and Sons.
Figure 21. TEM image (left) and particle size distribution (right) of the MPF-PPI-G3-Pd catalyst [362,390]. Reprinted with permission of John Wiley and Sons, from: Boronoev, M. P.; Zolotukhina, A. V.; Ignatyeva, V. I.; Terenina, M. V.; Maximov, A. L.; Karakhanov, E. A. Palladium Catalysts Based on Mesoporous Organic Materials in Semihydrogenation of Alkynes. Macromol. Symp. 2016, 363, 57–63 [390]. Copyright 2016, John Wiley and Sons.
Polymers 14 00981 g021
Figure 22. The synthesis of dendronized alternating block-co-polymer via double-click alkyne-azide and Diels-Alder cycloaddition [75,392]. Reprinted with permission of John Wiley and Sons, from: Tonga, M.; Yesilbag Tonga, G.; Seber, G.; Gok, O.; Sanyal, A. Dendronized Polystyrene via Orthogonal Double-Click Reactions. J. Polym. Sci. A: Polym. Chem. 2013, 51, 5029–5037 [392]. Copyright 2013, John Wiley and Sons.
Figure 22. The synthesis of dendronized alternating block-co-polymer via double-click alkyne-azide and Diels-Alder cycloaddition [75,392]. Reprinted with permission of John Wiley and Sons, from: Tonga, M.; Yesilbag Tonga, G.; Seber, G.; Gok, O.; Sanyal, A. Dendronized Polystyrene via Orthogonal Double-Click Reactions. J. Polym. Sci. A: Polym. Chem. 2013, 51, 5029–5037 [392]. Copyright 2013, John Wiley and Sons.
Polymers 14 00981 g022
Scheme 41. Covalent immobilization of PAMAM dendrimers on silica and subsequent post-modification with chiral capping [92].
Scheme 41. Covalent immobilization of PAMAM dendrimers on silica and subsequent post-modification with chiral capping [92].
Polymers 14 00981 sch041
Scheme 42. Post-immobilization of PPI dendrimers in the SBA-15 channels via urethane bridge interactions [74].
Scheme 42. Post-immobilization of PPI dendrimers in the SBA-15 channels via urethane bridge interactions [74].
Polymers 14 00981 sch042
Scheme 43. Convergent grafting of phosphine-terminated PETIM dendrimer to silica and subsequent Pd nanoparticle and complex formation [102].
Scheme 43. Convergent grafting of phosphine-terminated PETIM dendrimer to silica and subsequent Pd nanoparticle and complex formation [102].
Polymers 14 00981 sch043
Scheme 44. Immobilization of PPI dendrimers in the silica-polyamine composite and subsequent Pd nanoparticle deposition [167].
Scheme 44. Immobilization of PPI dendrimers in the silica-polyamine composite and subsequent Pd nanoparticle deposition [167].
Polymers 14 00981 sch044
Scheme 45. Selectivities in the hydrogenation of phenylacetylene and conjugated dienes at conversions higher than 90% in the presence of SPC-PPI-G3-Pd catalyst [171].
Scheme 45. Selectivities in the hydrogenation of phenylacetylene and conjugated dienes at conversions higher than 90% in the presence of SPC-PPI-G3-Pd catalyst [171].
Polymers 14 00981 sch045
Scheme 46. The synthesis of hybrid SiO2-PPI-G3 material [396].
Scheme 46. The synthesis of hybrid SiO2-PPI-G3 material [396].
Polymers 14 00981 sch046
Figure 23. TEM images of SiO2-PPI-G3 (a) and meso-SiO2-PPI-G2 materials (b) [396].
Figure 23. TEM images of SiO2-PPI-G3 (a) and meso-SiO2-PPI-G2 materials (b) [396].
Polymers 14 00981 g023
Scheme 47. The synthesis of hybrid meso-SiO2-PPI-G2 material [396].
Scheme 47. The synthesis of hybrid meso-SiO2-PPI-G2 material [396].
Polymers 14 00981 sch047
Scheme 48. The synthesis of hybrid meso-PPI-G2-Si material [193].
Scheme 48. The synthesis of hybrid meso-PPI-G2-Si material [193].
Polymers 14 00981 sch048
Figure 24. TEM images of catalysts: (a) meso-SiO2-PPI-G2-Ru and (b) SiO2-PPI-G3-Ru [396].
Figure 24. TEM images of catalysts: (a) meso-SiO2-PPI-G2-Ru and (b) SiO2-PPI-G3-Ru [396].
Polymers 14 00981 g024
Figure 25. TEM images of catalysts: (a) meso-SiO2-PPI-G2-Pd and (b) SiO2-PPI-G3-Pd [193].
Figure 25. TEM images of catalysts: (a) meso-SiO2-PPI-G2-Pd and (b) SiO2-PPI-G3-Pd [193].
Polymers 14 00981 g025
Figure 26. Hydrogenation of styrenes in the presence of dendrimer-based organo-silica palladium nanocatalysts [193].
Figure 26. Hydrogenation of styrenes in the presence of dendrimer-based organo-silica palladium nanocatalysts [193].
Polymers 14 00981 g026
Figure 27. Hydrogenation of alkynes (a) and dienes (b) in the presence of dendrimer-based organo-silica palladium nanocatalysts [193].
Figure 27. Hydrogenation of alkynes (a) and dienes (b) in the presence of dendrimer-based organo-silica palladium nanocatalysts [193].
Polymers 14 00981 g027
Scheme 49. Possible reaction pathways for hydrogenation of levulinic acid and its esters [401]. Reprinted with permission of John Wiley and Sons, from: Maximov, A. L.; Zolotukhina, A. V.; Mamedli, A. A.; Kulikov, L. A.; Karakhanov, E. A. Selective Levulinic Acid Hydrogenation in the Presence of Hybrid Dendrimer-Based Catalysts. Part I: Monometallic. ChemCatChem 2018, 10, 222–233 [401]. Copyright 2017, John Wiley and Sons.
Scheme 49. Possible reaction pathways for hydrogenation of levulinic acid and its esters [401]. Reprinted with permission of John Wiley and Sons, from: Maximov, A. L.; Zolotukhina, A. V.; Mamedli, A. A.; Kulikov, L. A.; Karakhanov, E. A. Selective Levulinic Acid Hydrogenation in the Presence of Hybrid Dendrimer-Based Catalysts. Part I: Monometallic. ChemCatChem 2018, 10, 222–233 [401]. Copyright 2017, John Wiley and Sons.
Polymers 14 00981 sch049
Scheme 50. Hydrogenation of levulinic acid in the presence of PPI dendrimer-stabilized or dendrimer-encapsulated Ru nanoparticles in water medium [401]. Reprinted with permission of John Wiley and Sons, from: Maximov, A. L.; Zolotukhina, A. V.; Mamedli, A. A.; Kulikov, L. A.; Karakhanov, E. A. Selective Levulinic Acid Hydrogenation in the Presence of Hybrid Dendrimer-Based Catalysts. Part I: Monometallic. ChemCatChem 2018, 10, 222–233 [401]. Copyright 2017, John Wiley and Sons.
Scheme 50. Hydrogenation of levulinic acid in the presence of PPI dendrimer-stabilized or dendrimer-encapsulated Ru nanoparticles in water medium [401]. Reprinted with permission of John Wiley and Sons, from: Maximov, A. L.; Zolotukhina, A. V.; Mamedli, A. A.; Kulikov, L. A.; Karakhanov, E. A. Selective Levulinic Acid Hydrogenation in the Presence of Hybrid Dendrimer-Based Catalysts. Part I: Monometallic. ChemCatChem 2018, 10, 222–233 [401]. Copyright 2017, John Wiley and Sons.
Polymers 14 00981 sch050
Figure 28. Hydrogenation of phenols (a) and levulinic acid and its esters (b) in the presence of dendrimer-based organo-silica ruthenium nanocatalysts [396,401]. Reprinted with permission from: Maximov, A. L.; Zolotukhina, A. V.; Mamedli, A. A.; Kulikov, L. A.; Karakhanov, E. A. Selective Levulinic Acid Hydrogenation in the Presence of Hybrid Dendrimer-Based Catalysts. Part I: Monometallic. ChemCatChem 2018, 10, 222–233 [401]. Copyright 2017, John Wiley and Sons.
Figure 28. Hydrogenation of phenols (a) and levulinic acid and its esters (b) in the presence of dendrimer-based organo-silica ruthenium nanocatalysts [396,401]. Reprinted with permission from: Maximov, A. L.; Zolotukhina, A. V.; Mamedli, A. A.; Kulikov, L. A.; Karakhanov, E. A. Selective Levulinic Acid Hydrogenation in the Presence of Hybrid Dendrimer-Based Catalysts. Part I: Monometallic. ChemCatChem 2018, 10, 222–233 [401]. Copyright 2017, John Wiley and Sons.
Polymers 14 00981 g028
Figure 29. TEM images of heterogeneous coordination dendritic catalyst PdII/CuII/DAB(CN)16 formed in situ [124]. Reprinted with permission of Elsevier, from: Karakhanov, E. A.; Maximov, A. L.; Tarasevich, B. N.; Skorkin, V. A. Dendrimer-based catalysts in Wacker-oxidation: Unexpected selectivity to terminal double bonds. J. Mol. Catal. A: Chem. 2009, 297, 73–79 [124]. Copyright 2008, Elsevier.
Figure 29. TEM images of heterogeneous coordination dendritic catalyst PdII/CuII/DAB(CN)16 formed in situ [124]. Reprinted with permission of Elsevier, from: Karakhanov, E. A.; Maximov, A. L.; Tarasevich, B. N.; Skorkin, V. A. Dendrimer-based catalysts in Wacker-oxidation: Unexpected selectivity to terminal double bonds. J. Mol. Catal. A: Chem. 2009, 297, 73–79 [124]. Copyright 2008, Elsevier.
Polymers 14 00981 g029
Figure 30. Formation of dendritic coordination networks and selective Wacker-oxidation of terminal alkenes in their presence [124]. Adapted with permission of Elsevier, from: Karakhanov, E.A.; Maximov, A. L.; Tarasevich, B. N.; Skorkin, V.A. Dendrimer-based catalysts in Wacker-oxidation: Unexpected selectivity to terminal double bonds. J. Mol. Catal. A: Chem. 2009, 297, 73–79 [124]. Copyright 2008, Elsevier.
Figure 30. Formation of dendritic coordination networks and selective Wacker-oxidation of terminal alkenes in their presence [124]. Adapted with permission of Elsevier, from: Karakhanov, E.A.; Maximov, A. L.; Tarasevich, B. N.; Skorkin, V.A. Dendrimer-based catalysts in Wacker-oxidation: Unexpected selectivity to terminal double bonds. J. Mol. Catal. A: Chem. 2009, 297, 73–79 [124]. Copyright 2008, Elsevier.
Polymers 14 00981 g030
Scheme 51. The synthesis of networks, based on PPI dendrimers, cross-linked with Sc3+ ions [404].
Scheme 51. The synthesis of networks, based on PPI dendrimers, cross-linked with Sc3+ ions [404].
Polymers 14 00981 sch051
Scheme 52. Mikayama and Diels-Alder reactions in the presence of network catalysts, based on PPI dendrimers, cross-linked with Sc3+ ions [404].
Scheme 52. Mikayama and Diels-Alder reactions in the presence of network catalysts, based on PPI dendrimers, cross-linked with Sc3+ ions [404].
Polymers 14 00981 sch052
Scheme 53. Modification of aryl phosphite dendrimers and subsequent synthesis of Pd complex catalyst precursor [405].
Scheme 53. Modification of aryl phosphite dendrimers and subsequent synthesis of Pd complex catalyst precursor [405].
Polymers 14 00981 sch053
Figure 31. A speculative representation aryl phosphite dendrimer binding to the coordination polymeric ensembles through the interaction of Pd (0) species or clusters with 15-membered triolefinic triaza-macrocycles on the dendrimer periphery [405].
Figure 31. A speculative representation aryl phosphite dendrimer binding to the coordination polymeric ensembles through the interaction of Pd (0) species or clusters with 15-membered triolefinic triaza-macrocycles on the dendrimer periphery [405].
Polymers 14 00981 g031
Figure 32. TEM images of Pd nanocatalysts, based on Caminade/Majoral dendrimers of the 0th generation, terminated by 15-membered triolefinic triaza-macrocycles [405]. Reprinted with permission of American Chemical Society, from: Badetti, E.; Caminade, A.-M.; Majoral, J.-P.; Moreno-Mañas, M.; Sebastián, R. M. Palladium (0) Nanoparticles Stabilized by Phosphorus Dendrimers Containing Coordinating 15-Membered Triolefinic Macrocycles in Periphery. Langmuir 2008, 24, 2090–2101 [405]. Copyright 2008, American Chemical Society.
Figure 32. TEM images of Pd nanocatalysts, based on Caminade/Majoral dendrimers of the 0th generation, terminated by 15-membered triolefinic triaza-macrocycles [405]. Reprinted with permission of American Chemical Society, from: Badetti, E.; Caminade, A.-M.; Majoral, J.-P.; Moreno-Mañas, M.; Sebastián, R. M. Palladium (0) Nanoparticles Stabilized by Phosphorus Dendrimers Containing Coordinating 15-Membered Triolefinic Macrocycles in Periphery. Langmuir 2008, 24, 2090–2101 [405]. Copyright 2008, American Chemical Society.
Polymers 14 00981 g032
Scheme 54. The synthesis of nanocatalysts based on the dendrimer networks, cross-linked with organic bifunctional agents [193,241,287].
Scheme 54. The synthesis of nanocatalysts based on the dendrimer networks, cross-linked with organic bifunctional agents [193,241,287].
Polymers 14 00981 sch054
Figure 33. Dendrimers and linkers used for dendritic network synthesis [122,123,127,128,241,287,408,409,412,413,414,415]. Reprinted and adapted with permissions of Elsevier, from: Karakhanov, E. A.; Maximov, A. L.; Zakharyan, E. M.; Zolotukhina, A. V.; Ivanov, A. O. Palladium nanoparticles on dendrimer-containing supports as catalysts for hydrogenation of unsaturated hydrocarbons. Mol. Catal. 2017, 440, 107–119 [128]; Karakhanov, E. A.; Maximov, A. L.; Zolotukhina, A. V. Selective semi-hydrogenation of phenyl acetylene by Pd nanocatalysts encapsulated into dendrimer networks. Mol. Catal. 2019, 469, 98–110 [241]. Copyright Elsevier, 2017 and 2019; and of Springer Nature, from: Karakhanov, E.; Maximov, A.; Zolotukhina, A.; Kardasheva, Yu.; Talanova, M. Thermo-responsive Ruthenium Dendrimer-based Catalysts for Hydrogenation of the Aromatic Compounds and Phenols. J. Inorg. Organomet. Polym. Mater. 2016, 26, 1264–1279 [127]. Copyright Springer, 2016.
Figure 33. Dendrimers and linkers used for dendritic network synthesis [122,123,127,128,241,287,408,409,412,413,414,415]. Reprinted and adapted with permissions of Elsevier, from: Karakhanov, E. A.; Maximov, A. L.; Zakharyan, E. M.; Zolotukhina, A. V.; Ivanov, A. O. Palladium nanoparticles on dendrimer-containing supports as catalysts for hydrogenation of unsaturated hydrocarbons. Mol. Catal. 2017, 440, 107–119 [128]; Karakhanov, E. A.; Maximov, A. L.; Zolotukhina, A. V. Selective semi-hydrogenation of phenyl acetylene by Pd nanocatalysts encapsulated into dendrimer networks. Mol. Catal. 2019, 469, 98–110 [241]. Copyright Elsevier, 2017 and 2019; and of Springer Nature, from: Karakhanov, E.; Maximov, A.; Zolotukhina, A.; Kardasheva, Yu.; Talanova, M. Thermo-responsive Ruthenium Dendrimer-based Catalysts for Hydrogenation of the Aromatic Compounds and Phenols. J. Inorg. Organomet. Polym. Mater. 2016, 26, 1264–1279 [127]. Copyright Springer, 2016.
Polymers 14 00981 g033
Figure 34. Hydrogenation of phenols in the presence of dendrimer network catalysts. Reaction conditions are: 85 °C, H2 30 atm., H2O [122,412].
Figure 34. Hydrogenation of phenols in the presence of dendrimer network catalysts. Reaction conditions are: 85 °C, H2 30 atm., H2O [122,412].
Polymers 14 00981 g034
Figure 35. Hydrogenation of unsaturated compounds in the presence of dendrimer network catalysts. Reaction conditions are: 80 °C, H2 5–10 atm. [128,241,409].
Figure 35. Hydrogenation of unsaturated compounds in the presence of dendrimer network catalysts. Reaction conditions are: 80 °C, H2 5–10 atm. [128,241,409].
Polymers 14 00981 g035
Scheme 55. The synthesis of “mesoporous” dendrimer network catalysts, using template method, and the activities reached in the hydrogenation compounds [128]. Reprinted with permission of Elsevier, from: Karakhanov, E. A.; Maximov, A. L.; Zakharyan, E. M.; Zolotukhina, A. V.; Ivanov, A. O. Palladium nanoparticles on dendrimer-containing supports as catalysts for hydrogenation of unsaturated hydrocarbons. Mol. Catal. 2017, 440, 107–119 [128]. Copyright 2017, Elsevier.
Scheme 55. The synthesis of “mesoporous” dendrimer network catalysts, using template method, and the activities reached in the hydrogenation compounds [128]. Reprinted with permission of Elsevier, from: Karakhanov, E. A.; Maximov, A. L.; Zakharyan, E. M.; Zolotukhina, A. V.; Ivanov, A. O. Palladium nanoparticles on dendrimer-containing supports as catalysts for hydrogenation of unsaturated hydrocarbons. Mol. Catal. 2017, 440, 107–119 [128]. Copyright 2017, Elsevier.
Polymers 14 00981 sch055
Figure 36. Phenylacetylene hydrogenation at different temperatures in the presence of DAB-PPI-G3-DMDPDI-Pd catalyst. Reaction conditions are S/Pd = 12,930, H2 30 atm. [128]. Reprinted with permission of Elsevier, from: from Karakhanov, E. A.; Maximov, A. L.; Zakharyan, E. M.; Zolotukhina, A. V.; Ivanov, A. O. Palladium nanoparticles on dendrimer-containing supports as catalysts for hydrogenation of unsaturated hydrocarbons. Mol. Catal. 2017, 440, 107–119 [128]. Copyright 2017, Elsevier.
Figure 36. Phenylacetylene hydrogenation at different temperatures in the presence of DAB-PPI-G3-DMDPDI-Pd catalyst. Reaction conditions are S/Pd = 12,930, H2 30 atm. [128]. Reprinted with permission of Elsevier, from: from Karakhanov, E. A.; Maximov, A. L.; Zakharyan, E. M.; Zolotukhina, A. V.; Ivanov, A. O. Palladium nanoparticles on dendrimer-containing supports as catalysts for hydrogenation of unsaturated hydrocarbons. Mol. Catal. 2017, 440, 107–119 [128]. Copyright 2017, Elsevier.
Polymers 14 00981 g036
Figure 37. Hydrogenation of aromatic compounds in the presence of catalysts, based on PPI dendrimer networks. Reaction conditions are 80 °C, H2O, H2 30 atm. [408].
Figure 37. Hydrogenation of aromatic compounds in the presence of catalysts, based on PPI dendrimer networks. Reaction conditions are 80 °C, H2O, H2 30 atm. [408].
Polymers 14 00981 g037
Scheme 56. The synthesis of PAMAM dendrimer-encapsulated Pd nanoparticles, in situ immobilized in the cross-linked polymer matrix [26,417]. Reprinted by permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A.V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Scheme 56. The synthesis of PAMAM dendrimer-encapsulated Pd nanoparticles, in situ immobilized in the cross-linked polymer matrix [26,417]. Reprinted by permission of Springer Nature, from: Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A.V.; Kardasheva, Yu. S. Hydrogenation catalysts based on metal nanoparticles stabilized by organic ligand. Russ. Chem. Bull. Int. Ed. 2013, 62, 1465–1492 [26]. Copyright 2014, Springer.
Polymers 14 00981 sch056
Scheme 57. The synthesis of PEI dendrimer-encapsulated Pd nanoparticles, in situ immobilized in the resorcinol-formaldehyde polymer network [416].
Scheme 57. The synthesis of PEI dendrimer-encapsulated Pd nanoparticles, in situ immobilized in the resorcinol-formaldehyde polymer network [416].
Polymers 14 00981 sch057
Figure 38. Dendrimer-encapsulated (left) and dendrimer core magnetically active nanoparticles (right), which may be used as homogeneous catalysts or their precursors [62,419,424,425].
Figure 38. Dendrimer-encapsulated (left) and dendrimer core magnetically active nanoparticles (right), which may be used as homogeneous catalysts or their precursors [62,419,424,425].
Polymers 14 00981 g038
Figure 39. The typical structure of heterogeneous magnetically separable dendrimer-based catalyst in the example of iron oxide nanoparticle-cored melamine dendrimer [419,425].
Figure 39. The typical structure of heterogeneous magnetically separable dendrimer-based catalyst in the example of iron oxide nanoparticle-cored melamine dendrimer [419,425].
Polymers 14 00981 g039
Scheme 58. The synthesis of Pd nanoparticles, encapsulated in polyphenylenepyridyl dendrons, covalently attached to magnetic Fe3O4 core [426,427,428].
Scheme 58. The synthesis of Pd nanoparticles, encapsulated in polyphenylenepyridyl dendrons, covalently attached to magnetic Fe3O4 core [426,427,428].
Polymers 14 00981 sch058
Figure 40. TEM image of Den-Fe3O4 magnetic-sensitive dendronized carrier (left) [428] and HRTEM image of Pd-Den-Fe3O4 catalyst (right) [426]. Reprinted by permission of Royal Society of Chemistry, from: Yuzik-Klimova, E. Y.; Kuchkina, N. V.; Sorokina, S. A.; Morgan, D. G.; Boris, B.; Nikoshvili, L.Zh.; Lyubimova, N. A.; Matveeva, V. G.; Sulman, E. M.; Stein, B. D.; Mahmoud, W. E.; Al-Ghamdi, A. A.; Kostopoulou, A.; Lappas, A.; Shifrina, Z.B.; Bronstein, L. M. Magnetically Recoverable Catalysts Based on Polyphenylenepyridyl Dendrons and Dendrimers. RSC Adv. 2014, 4, 23271–23280 [428]. Copyright 2014, Royal Society of Chemistry.
Figure 40. TEM image of Den-Fe3O4 magnetic-sensitive dendronized carrier (left) [428] and HRTEM image of Pd-Den-Fe3O4 catalyst (right) [426]. Reprinted by permission of Royal Society of Chemistry, from: Yuzik-Klimova, E. Y.; Kuchkina, N. V.; Sorokina, S. A.; Morgan, D. G.; Boris, B.; Nikoshvili, L.Zh.; Lyubimova, N. A.; Matveeva, V. G.; Sulman, E. M.; Stein, B. D.; Mahmoud, W. E.; Al-Ghamdi, A. A.; Kostopoulou, A.; Lappas, A.; Shifrina, Z.B.; Bronstein, L. M. Magnetically Recoverable Catalysts Based on Polyphenylenepyridyl Dendrons and Dendrimers. RSC Adv. 2014, 4, 23271–23280 [428]. Copyright 2014, Royal Society of Chemistry.
Polymers 14 00981 g040
Scheme 59. Reactions, catalyzed by Pd-Den-Fe3O4 nanoparticles [426,427,428,429].
Scheme 59. Reactions, catalyzed by Pd-Den-Fe3O4 nanoparticles [426,427,428,429].
Polymers 14 00981 sch059
Figure 41. Palladium complex catalyst, based on poly(alkyl aryl ether) dendrons, covalently attached to magnetic Fe2O3 core coated with carboxylic polymer [430]. Adapted with permission of John Wiley and Sons, from: Rosario-Amorin, D.; Gaboyard, M.; Clérac, R.; Vellutini, L.; Nlate, S.; Heuzé, K. Metallodendritic Grafted Core–Shell γ-Fe2O3 Nanoparticles Used as Recoverable Catalysts in Suzuki C–C Coupling Reactions. Chem. Eur. J. 2012, 18, 3305–3315 [430]. Copyright John Wiley and Sons, 2012.
Figure 41. Palladium complex catalyst, based on poly(alkyl aryl ether) dendrons, covalently attached to magnetic Fe2O3 core coated with carboxylic polymer [430]. Adapted with permission of John Wiley and Sons, from: Rosario-Amorin, D.; Gaboyard, M.; Clérac, R.; Vellutini, L.; Nlate, S.; Heuzé, K. Metallodendritic Grafted Core–Shell γ-Fe2O3 Nanoparticles Used as Recoverable Catalysts in Suzuki C–C Coupling Reactions. Chem. Eur. J. 2012, 18, 3305–3315 [430]. Copyright John Wiley and Sons, 2012.
Polymers 14 00981 g041
Scheme 60. Suzuki cross-coupling in the presence of heterogeneous dendronized magnetic nanocatalysts [430].
Scheme 60. Suzuki cross-coupling in the presence of heterogeneous dendronized magnetic nanocatalysts [430].
Polymers 14 00981 sch060
Figure 42. PEG-terminated poly(carbosilane/triazol/aryl ether) dendrons, covalently attached to γ-Fe2O3 superparamagnetic nanoparticle coated with silica [431,432]. Reprinted and adapted with permission of John Wiley and Sons, from: Wang, D.; Deraedt, C.; Salmon, L.; Labrugère, C.; Etienne, L.; Ruiz, J.; Astruc D. Efficient and Magnetically Recoverable “Click” PEGylated γ-Fe2O3–Pd Nanoparticle Catalysts for Suzuki–Miyaura, Sonogashira, and Heck Reactions with Positive Dendritic Effects. Chem. Eur. J. 2015, 21, 1508−1519 [432]. Copyright 2014, John Wiley and Sons.
Figure 42. PEG-terminated poly(carbosilane/triazol/aryl ether) dendrons, covalently attached to γ-Fe2O3 superparamagnetic nanoparticle coated with silica [431,432]. Reprinted and adapted with permission of John Wiley and Sons, from: Wang, D.; Deraedt, C.; Salmon, L.; Labrugère, C.; Etienne, L.; Ruiz, J.; Astruc D. Efficient and Magnetically Recoverable “Click” PEGylated γ-Fe2O3–Pd Nanoparticle Catalysts for Suzuki–Miyaura, Sonogashira, and Heck Reactions with Positive Dendritic Effects. Chem. Eur. J. 2015, 21, 1508−1519 [432]. Copyright 2014, John Wiley and Sons.
Polymers 14 00981 g042
Scheme 61. The synthesis of Pd nanoparticles, stabilized by PEG-terminated triazol arene-cored dendrimers, immobilized on silica-coated γ-Fe2O3 superparamagnetic core by means of hydrogen bonds [431,433]. Adapted with permission of John Wiley and Sons, from: Deraedt, C.; Wang, D.; Salmon, L.; Etienne, L.; Labrugère, C.; Ruiz, J.; Astruc, D. Robust, efficient, and recyclable catalyst by impregnation of dendritically preformed Pd nanoparticles on magnetic support. ChemCatChem 2015, 7, 303–308 [433]. Copyright 2014, John Wiley and Sons.
Scheme 61. The synthesis of Pd nanoparticles, stabilized by PEG-terminated triazol arene-cored dendrimers, immobilized on silica-coated γ-Fe2O3 superparamagnetic core by means of hydrogen bonds [431,433]. Adapted with permission of John Wiley and Sons, from: Deraedt, C.; Wang, D.; Salmon, L.; Etienne, L.; Labrugère, C.; Ruiz, J.; Astruc, D. Robust, efficient, and recyclable catalyst by impregnation of dendritically preformed Pd nanoparticles on magnetic support. ChemCatChem 2015, 7, 303–308 [433]. Copyright 2014, John Wiley and Sons.
Polymers 14 00981 sch061
Scheme 62. Reactions, catalyzed by Pd nanoparticles, immobilized on γ-Fe2O3/SiO2 core support, coated with PEGylated poly(carbosilane/triazol/aryl ether) dendrimers and dendrons [10,432,433].
Scheme 62. Reactions, catalyzed by Pd nanoparticles, immobilized on γ-Fe2O3/SiO2 core support, coated with PEGylated poly(carbosilane/triazol/aryl ether) dendrimers and dendrons [10,432,433].
Polymers 14 00981 sch062
Figure 43. Rhodium complex catalyst, based on PAMAM dendrons, covalently attached to silica-coated magnetic Fe3O4 core [434]. Reprinted (adapted) with permission of American Chemical Society, from: Abu-Reziq, R.; Alper, H.; Wang, D.; Post, M. L. Metal Supported on Dendronized Magnetic Nanoparticles: Highly Selective Hydroformylation Catalysts. J. Am. Chem. Soc. 2006, 128, 5279–5282 [434]. Copyright 2006, American Chemical Society.
Figure 43. Rhodium complex catalyst, based on PAMAM dendrons, covalently attached to silica-coated magnetic Fe3O4 core [434]. Reprinted (adapted) with permission of American Chemical Society, from: Abu-Reziq, R.; Alper, H.; Wang, D.; Post, M. L. Metal Supported on Dendronized Magnetic Nanoparticles: Highly Selective Hydroformylation Catalysts. J. Am. Chem. Soc. 2006, 128, 5279–5282 [434]. Copyright 2006, American Chemical Society.
Polymers 14 00981 g043
Figure 44. Heterogeneous magnetically recoverable Os-containing catalysts for alkene dihydroxylation: control non-dendritic catalyst (left) and dendron-doped (right) [435].
Figure 44. Heterogeneous magnetically recoverable Os-containing catalysts for alkene dihydroxylation: control non-dendritic catalyst (left) and dendron-doped (right) [435].
Polymers 14 00981 g044
Scheme 63. Dihydroxylation of olefins in the presence of Os catalysts [435].
Scheme 63. Dihydroxylation of olefins in the presence of Os catalysts [435].
Polymers 14 00981 sch063
Scheme 64. Thermo-responsive release and catch system for effective magnet-assisted dendritic catalyst recycling [419,436].
Scheme 64. Thermo-responsive release and catch system for effective magnet-assisted dendritic catalyst recycling [419,436].
Polymers 14 00981 sch064
Scheme 65. Suzuki-cross-coupling reaction in the presence of pyrene-tagged polyphosphazene dendrimers and Co magnetic nanoparticles [436].
Scheme 65. Suzuki-cross-coupling reaction in the presence of pyrene-tagged polyphosphazene dendrimers and Co magnetic nanoparticles [436].
Polymers 14 00981 sch065
Table 1. Application of metal complex and nanoparticle dendrimer-based material in catalytic processes.
Table 1. Application of metal complex and nanoparticle dendrimer-based material in catalytic processes.
EntryProcessMetal Complex Catalysts [Refs.]Nanoparticle Catalysts [Refs.]
1Hydrogenation of alkenes[119,120,139,214,215,216,217,218][116,117,123,128,129,137,150,152,161,162,163,165,166,178,219,220]
2Hydrogenation of phenols, aromatic, and heterocyclic compounds[221][122,127]
3Semihydrogenation of alkynes, dienes, and polyenes[140,148][116,123,128,131,150,156,159,160,164,171,193,207]
4Hydrogenation and hydroamination of carbonyl compounds[222][179,203,217,223]
5Nitroarene reduction [166,169,170,186,187,191,192,200,202,203,219,224,225,226,227,228,229]
6Oxygen reduction [153,230]
7Hydroformylation[146,217,231]
8Alkyne-azide cycloaddition[188,190,232][233]
9Allylic addition[132,234,235,236,237,238][157]
10Asymmetric synthesis and catalysis [238,239,240]
11Wacker oxidation[124,147,241]
12C–H, O–H and sulfide oxidation[138,242,243,244][180,183,245,246]
13Epoxidation[247,248,249,250,251,252,253]
14Oligomerization and polymerization[254,255,256,257,258,259,260,261,262,263,264,265,266]
15Metathesis[264,265,266,267,268,269,270]
16Oxidative coupling[109,271,272,273]
17Cross-coupling[101,108,130,132,142,143,144,274][118,121,125,126,129,186,189,275,276,277,278]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Karakhanov, E.; Maximov, A.; Zolotukhina, A. Heterogeneous Dendrimer-Based Catalysts. Polymers 2022, 14, 981. https://doi.org/10.3390/polym14050981

AMA Style

Karakhanov E, Maximov A, Zolotukhina A. Heterogeneous Dendrimer-Based Catalysts. Polymers. 2022; 14(5):981. https://doi.org/10.3390/polym14050981

Chicago/Turabian Style

Karakhanov, Eduard, Anton Maximov, and Anna Zolotukhina. 2022. "Heterogeneous Dendrimer-Based Catalysts" Polymers 14, no. 5: 981. https://doi.org/10.3390/polym14050981

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop