chem soc rev - imperial college london · 2018. 7. 10. · borane), this means that flps are...

12
Chem Soc Rev TUTORIAL REVIEW This journal is © The Royal Society of Chemistry 2017 Chem. Soc . Rev., 2017, 00, 1-3 | 1 Please do not adjust margins Please do not adjust margins Department of Chemistry, Imperial College London, SW7 2AZ, UK. E-mail: [email protected] Received 00th January 20xx, Accepted 00th January 20xx DOI: 10.1039/x0xx00000x www.rsc.org/ Designing Effective ‘Frustrated Lewis Pair’ Hydrogenation Catalysts Daniel J. Scott,* Matthew J. Fuchter and Andrew E. Ashley* The past decade has seen the subject of transition metal-free catalytic hydrogenation develop incredibly rapidly, transforming from a largely hypothetical possibility to a well-established field that can be applied to the reduction of a diverse variety of functional groups under mild conditions. This remarkable change is principally attributable to the development of so-called ‘frustrated Lewis pairs’: unquenched combinations of bulky Lewis acids and bases whose dual reactivity can be exploited for the facile activation of otherwise inert chemical bonds. While a number of comprehensive reviews into frustrated Lewis pair chemistry have been published in recent years, this tutorial review aims to provide a focused guide to the development of efficient FLP hydrogenation catalysts, through identification and consideration of the key factors that govern their effectiveness. Following discussion of these factors, their importance will be illustrated using a case study from our own research, namely the development of FLP protocols for successful hydrogenation of aldehydes and ketones, and for related moisture-tolerant hydrogenation. Introduction to FLP chemistry Since the earliest days of the field, the study of homogeneous catalysis has been all but synonymous with the study of transition metal (TM) catalysis, particularly in the activation of relatively inert small molecules or of strong chemical bonds. The privileged reactivity demonstrated by TM compounds can be attributed to their characteristic electronic structures, with partially occupied sets of d-orbitals leading to the simultaneous presence of both nucleophilic/Lewis basic and electrophilic/Lewis acidic frontier orbitals located on the same atom. It is the ability of both types of orbital to interact synergistically with a substrate that allows for the activation of functional groups that would normally be kinetically inert, even where these groups would be unreactive towards a Lewis acidic or Lewis basic site on its own (illustrated for H 2 in Fig. 1a). Comparable electronic structures are uncommon for stable main group compounds, which explains their general inability to mediate similar catalytic reactions. Nevertheless, some examples do exist (notably the various well-known carbenes and related group IV R 2 E species) and in recent years there has been great interest in the isolation and study of such compounds in the hope that they may demonstrate a similar potential for catalysis, and ultimately provide counterparts or alternatives to TMs (many of which suffer from high toxicity, high cost, or low abundance). 1 Indeed, some such compounds have been shown to readily undergo a variety of ‘TM-like’ reactions. For example, Bertrand et al. were able to demonstrate activation of inert E—H bonds (E = N, H) by addition to singlet carbenes, with the observed reactivity attributed to simultaneous interaction of the substrate with the electrophilic 2p and nucleophilic sp 2 orbitals on the reactive carbon centre, in a manner clearly reminiscent of TMs (Fig. 1b). 2 Nevertheless, the adaptation of stoichiometric bond activation chemistry by main group compounds into useful catalytic cycles has proven highly challenging, and only very few such examples have been reported. This can broadly be attributed to the typical low stability of unsaturated p-block compounds, which leads to difficulties in catalyst regeneration (and thus prevents closure of the catalytic cycle) and tendency towards decomposition, as well as more general difficulties in initial isolation and handling. Key learning points 1. Rational FLP design must be based on an understanding of the relevant key mechanistic steps. 2. H + and H affinities are crucial parameters and must be balanced relative to both the substrate and each one another. 3. Reactivity can be inhibited by either exceedingly high or low steric bulk, and the ideal profile will be substrate-dependent. 4. Intramolecular FLPs offer the possibility of improved reactivity, but at the cost of more challenging catalyst development. 5. Successful FLP design requires an understanding of inhibition/decomposition mechanisms, which are often LA-related.

Upload: others

Post on 13-Feb-2021

5 views

Category:

Documents


0 download

TRANSCRIPT

  • Chem Soc Rev

    TUTORIAL REVIEW

    This journal is © The Royal Society of Chemistry 2017 Chem. Soc. Rev., 2017, 00, 1-3 | 1

    Please do not adjust margins

    Please do not adjust margins

    Department of Chemistry, Imperial College London, SW7 2AZ, UK. E-mail: [email protected]

    Received 00th January 20xx,

    Accepted 00th January 20xx

    DOI: 10.1039/x0xx00000x

    www.rsc.org/

    Designing Effective ‘Frustrated Lewis Pair’ Hydrogenation Catalysts

    Daniel J. Scott,* Matthew J. Fuchter and Andrew E. Ashley*

    The past decade has seen the subject of transition metal-free catalytic hydrogenation develop incredibly rapidly,

    transforming from a largely hypothetical possibility to a well-established field that can be applied to the reduction of a

    diverse variety of functional groups under mild conditions. This remarkable change is principally attributable to the

    development of so-called ‘frustrated Lewis pairs’: unquenched combinations of bulky Lewis acids and bases whose dual

    reactivity can be exploited for the facile activation of otherwise inert chemical bonds. While a number of comprehensive

    reviews into frustrated Lewis pair chemistry have been published in recent years, this tutorial review aims to provide a

    focused guide to the development of efficient FLP hydrogenation catalysts, through identification and consideration of the

    key factors that govern their effectiveness. Following discussion of these factors, their importance will be illustrated using

    a case study from our own research, namely the development of FLP protocols for successful hydrogenation of aldehydes

    and ketones, and for related moisture-tolerant hydrogenation.

    Introduction to FLP chemistry

    Since the earliest days of the field, the study of homogeneous

    catalysis has been all but synonymous with the study of

    transition metal (TM) catalysis, particularly in the activation of

    relatively inert small molecules or of strong chemical bonds.

    The privileged reactivity demonstrated by TM compounds can

    be attributed to their characteristic electronic structures, with

    partially occupied sets of d-orbitals leading to the

    simultaneous presence of both nucleophilic/Lewis basic and

    electrophilic/Lewis acidic frontier orbitals located on the same

    atom. It is the ability of both types of orbital to interact

    synergistically with a substrate that allows for the activation of

    functional groups that would normally be kinetically inert,

    even where these groups would be unreactive towards a Lewis

    acidic or Lewis basic site on its own (illustrated for H2 in Fig.

    1a). Comparable electronic structures are uncommon for

    stable main group compounds, which explains their general

    inability to mediate similar catalytic reactions. Nevertheless,

    some examples do exist (notably the various well-known

    carbenes and related group IV R2E species) and in recent years

    there has been great interest in the isolation and study of such

    compounds in the hope that they may demonstrate a similar

    potential for catalysis, and ultimately provide counterparts or

    alternatives to TMs (many of which suffer from high toxicity,

    high cost, or low abundance).1 Indeed, some such compounds

    have been shown to readily undergo a variety of ‘TM-like’

    reactions. For example, Bertrand et al. were able to

    demonstrate activation of inert E—H bonds (E = N, H) by

    addition to singlet carbenes, with the observed reactivity

    attributed to simultaneous interaction of the substrate with

    the electrophilic 2p and nucleophilic sp2 orbitals on the

    reactive carbon centre, in a manner clearly reminiscent of TMs

    (Fig. 1b).2

    Nevertheless, the adaptation of stoichiometric bond

    activation chemistry by main group compounds into useful

    catalytic cycles has proven highly challenging, and only very

    few such examples have been reported. This can broadly be

    attributed to the typical low stability of unsaturated p-block

    compounds, which leads to difficulties in catalyst regeneration

    (and thus prevents closure of the catalytic cycle) and tendency

    towards decomposition, as well as more general difficulties in

    initial isolation and handling.

    Key learning points 1. Rational FLP design must be based on an understanding of the relevant key mechanistic steps. 2. H

    + and H

    – affinities are crucial parameters and must be balanced relative to both the substrate and each one another.

    3. Reactivity can be inhibited by either exceedingly high or low steric bulk, and the ideal profile will be substrate-dependent. 4. Intramolecular FLPs offer the possibility of improved reactivity, but at the cost of more challenging catalyst development. 5. Successful FLP design requires an understanding of inhibition/decomposition mechanisms, which are often LA-related.

  • ARTICLE Journal Name

    2 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

    Please do not adjust margins

    Please do not adjust margins

    In 2006 Stephan and co-workers described results that

    have led to an alternative and much simpler approach for

    obtaining TM-like reactivity using main group compounds.3

    The authors observed that under an atmosphere of H2 a

    solution of an intramolecular phosphine-borane was converted

    to into the zwitterionic phosphonium borohydride, via

    activation and cleavage of the homopolar H—H bond (Fig. 2a).

    It was quickly realised that this reactivity could be generalised

    to simple intermolecular phosphine/borane combinations,

    provided that their steric bulk was sufficient to prevent adduct

    formation between the Lewis acid (LA) and base (LB).4

    Because of their sterically-induced inability to quench one

    another, such systems have come to be known as ‘frustrated’

    Lewis pairs (FLPs).5 Subsequent work by many groups has

    shown that FLP reactivity can be observed with a much wider

    variety of both inter- and intramolecular LA and LB

    combinations [e.g. boranes, boreniums, alanes, carbocations,

    silyliums and stannyliums; phosphines, amines and N-

    heterocyclic carbenes (NHCs)], and can lead to activation of a

    great many other small molecules and chemical bonds (e.g.

    CO2 and other p-block oxides; alkenes and alkynes; acidic and

    hydridic E—H bonds).6

    The TM-like reactivity of FLPs has again been attributed to

    the simultaneous action on the H2 molecule of energetically-

    accessible Lewis acidic and basic orbitals (Fig. 1c).7 However,

    unlike the other examples discussed so far, in FLPs these

    orbitals are spatially separated from one another, and

    localised on different functional groups. As a consequence, it

    is typically relatively easy to fine-tune the properties of one

    (e.g. sterics or electronics) without having a significant impact

    on the other. Given also that FLPs are readily constructed from

    robust, well-understood functional groups (usually an amine or

    phosphine combined with a strong fluoroaryl-substituted

    borane), this means that FLPs are uniquely well-suited among

    unsaturated p-block compounds for the development of

    catalytic applications. Indeed, within two years of the first

    report of FLP H2 activation, the same authors also described

    the first example of FLP-catalysed hydrogenation; the

    conversion of simple imines to amines (Fig. 2b).8 Subsequent

    rapid progress has expanded the scope of FLP-catalysed

    hydrogenations to include substrates ranging from alkenes and

    aromatics to aldehydes and ketones.6 FLPs have thus provided

    the first general methodology for catalytic hydrogenation that

    does not require the use of a TM.

    Dr Matthew Fuchter is a Reader in

    Chemistry at Imperial College. The

    Fuchter group has a wide-ranging

    track record in the design, synthesis

    and application of organic molecules

    in chemistry, medicine and materials.

    Representative examples include the

    design and development of novel

    bioactive probes, the study of novel

    chiral semiconducting molecules, and

    the development of novel FLP catalysts.

    Daniel Scott is an EPSRC doctoral

    prize fellow currently working in the

    group of Dr Andrew E. Ashley at

    Imperial College, where he had

    previously obtained his PhD studying

    the development of FLP catalysis. His

    current research focuses on the

    development of Fe-based catalysts

    for homogeneous N2 fixation.

  • Journal Name ARTICLE

    This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 3

    Please do not adjust margins

    Please do not adjust margins

    Fig. 3 H2 activation by intramolecular and intermolecular FLPs. For

    the latter, the termolecular reaction step is facilitated by formation

    of a weakly-bound ‘encounter complex’ between the LA and LB,

    into which H2 can add.

    A note on FLPs and other branches of chemistry

    Given the simplicity of the FLP concept, it is perhaps not

    surprising that it has begun to be invoked in relation to quite a

    broad range of chemical processes. This includes discussion of

    newly developed or discovered reactions, but also of many

    that pre-date the FLP formalism. Notable examples in the

    latter category include Piers-type hydrosilylation,9 metal-ligand

    cooperative catalysis,10

    and the chemistry of solid surfaces,11

    among many others. In one particularly dramatic example,

    KOR (R = alkyl) was reported to catalyse ketone hydrogenation

    under very forcing conditions, with H2 activated by ‘the joint

    action of a […] base and a Lewis-acid […] on the H2 molecule’,

    clearly foreshadowing the development of FLP-catalysed

    hydrogenation.12

    An analogy can also be drawn between FLP

    H2 activation and the chemistry of TM·(H2) complexes, where

    binding to a Lewis acidic TM creates a Brønsted acidic H2

    moiety that can be deprotonated by LBs.13

    As a consequence,

    it can sometimes be unclear where the formal boundaries of

    the ‘FLP catalysis’ field should be placed (for example, where it

    may overlap with LA catalysis, especially in reactions that do

    not involve a co-catalytic LB). Ultimately, it is up to the

    individual chemist to decide whether invocation of the FLP

    concept is helpful in understanding the system in question, as

    is the case for many other descriptive models of chemistry

    (e.g. valence bond versus molecular orbital theory).

    Tutorial review aims and scope

    In this tutorial review we will outline the key factors that can

    determine the outcome of FLP-catalysed hydrogenation

    reactions, and illustrate how these principles can be used in

    the design of effective catalysts. Note that while they will not

    be discussed here explicitly, most of these principles will also

    be directly applicable to the development of various other FLP-

    catalysed reactions.

    Key aspects of FLP catalyst design

    Understanding the reaction mechanism

    As with any chemical transformation, the rational

    development of effective FLP hydrogenation catalysis is above

    all dependent upon a basic understanding of the key steps

    underlying the reaction mechanism. Understanding the

    mechanism by which FLPs are able to activate H2 is thus clearly

    central for considerations of hydrogenation catalysis. As

    already discussed, H—H cleavage is believed to occur through

    simultaneous interaction with both the LA and LB (Fig. 1c).

    While for intramolecular FLPs this leads to a feasible

    bimolecular reaction, for intermolecular systems it implies the

    need for an entropically unfavourable termolecular step. This

    in turn indicates that some transient interaction must form

    between two of the reaction components prior to the

    involvement of the third, in order to render the bond cleavage

    step kinetically accessible. At first it was supposed that this

    was likely to be either a weak LA←H2 or LB→H2 interaction.

    However, after initial attempts to find either experimental or

    computational evidence for such interactions were

    unsuccessful, further theoretical studies instead suggested

    formation of so-called ‘encounter complexes’ in which the LA

    and LB are held together by weak intermolecular interactions

    in such a way that they are pre-organised for subsequent H2

    activation (Fig. 3).14

    Subsequent experimental work has

    confirmed the existence of intramolecular interactions for two

    PR3/B(C6F5)3 FLPs (R = tBu, or Mes; 2,4,6-trimethylphenyl),

    through NMR spectroscopic techniques (e.g. observation of

    intermolecular 1H/

    19F correlations via 2D HOESY; Fig. 3), and as

    such this is believed to be the general mechanism by which H2

    activation is effected by a diverse range of FLPs.15

    Nevertheless, it should be emphasised that these

  • ARTICLE Journal Name

    4 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

    Please do not adjust margins

    Please do not adjust margins

    investigations have largely been limited to ‘typical’

    amine/borane or phosphine/borane FLPs, and alternative

    mechanisms cannot conclusively be ruled out in other

    intermolecular systems.

    Following H2 activation, catalytic hydrogenation requires

    transfer of the resulting H+ and H

    – fragments to the substrate

    in order to close the catalytic cycle. In almost all examples, this

    is believed to involve initial protonation of the substrate in

    order to activate it towards subsequent hydride transfer (Fig.

    4a). This can be attributed to the ubiquitous use in FLP

    hydrogenation catalysts of boranes incorporating very strongly

    electron-withdrawing fluoroaryl-substituents as LAs; following

    H2 activation these form relatively stable [Ar3BH]– anions that

    are not sufficiently powerful hydride donors to reduce the

    unactivated substrate. Only after protonation (or, at a

    minimum, hydrogen-bonding to [LB·H]+) does the substrate

    become sufficiently electrophilic for further reaction to occur.

    Nevertheless, some hydrogenations can proceed via

    alternative mechanisms, particularly where less ‘typical’ LAs

    are used. These may involve hydride transfer prior to

    protonation (if the substrate is sufficiently electrophilic and

    can stabilise the resulting negative charge), direct

    hydroelementation of the substrate, or activation of the

    substrate through coordination of the LA rather than H+ (Fig.

    4b). In some cases it is also possible that hydrogenation can

    proceed without the need to add an auxiliary LB catalyst (‘LA-

    only catalysis’), if the substrate is sufficiently basic and can act

    as the basic component to activate H2 directly in combination

    with the LA (Fig. 4c; imines, for example, are commonly

    hydrogenated by this mechanism). When attempting to

    rationally design or optimise a reaction it is crucial to

    determine if one of these alternative mechanisms might be

    operative. This can usually be achieved fairly simply through

    stoichiometric reaction of the substrate with pre-formed

    [LA·H]– reagents (e.g. [Bu4N]

    +[LA·H]

    – salts, which contain an

    inert countercation) in the presence and absence of possible

    activators (such as additional LA or ‘H+[WCA]

    –’, where [WCA]

    is a weakly-coordinating anion; Fig. 5). Likewise, computational

    studies may be used to provide additional insight.

    Tailoring LA and LB strength

    The ‘strength’ of the LA and LB components used to construct

    an FLP are of crucial importance to the success of FLP-

    catalysed hydrogenation reactions. FLP H2 activation has been

    reported using LBs that vary in strength by over 20 pKa units,

    and LAs whose calculated hydride ion affinities (G for LA+H–

    →[LA·H]–) vary by more than 140 kcal/mol. In an important

    study, Pápai and co-workers analysed the thermodynamics by

    FLP H2 activation by considering it as the sum of five separate

    conceptual steps (Fig. 6):16

    Heterolytic cleavage of H2 into H+ and H

    The size of this term will depend on factors such as the solvent

    (with more polar solvents making ionisation more favourable),

    but is independent of the FLP used.

  • Journal Name ARTICLE

    This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 5

    Please do not adjust margins

    Please do not adjust margins

    Separation of any LA←LB adduct into the uncoordinated Lewis

    pair

    For an FLP this term must, by definition, be close to zero. In

    some instances FLP reactivity can be observed for Lewis pairs

    that do form a weak adduct (discussed in the next section);

    however, even in these cases this term would not typically be

    expected to be too large.

    Attachment of H+ and H

    – to the LB and LA, respectively

    These terms are highly variable and depend above all on the

    choice of LA and LB.

    Any stabilising interaction between the resulting [LA·H]– and

    [LB·H]+ moieties (e.g. ion pairing, dihydrogen bonding)

    Computational studies have suggested that this term, while

    not negligible in magnitude, does not vary significantly across a

    selection of intermolecular FLPs (only neutral LAs and LBs were

    considered; it is not clear to what extent this conclusion will

    hold for charged species).16

    For intramolecular systems the

    [LA·H]–/[LB·H]

    + pairing term is typically larger, as enthalpically-

    favourable ion pairing can be achieved without such a

    significant entropic penalty. H2 activation in these systems is

    thus generally more favourable than in intermolecular systems

    of equivalent LA/LB strength. The magnitude of this additional

    stabilisation can be highly variable, and depends on the linker

    used (often unpredictably; vide infra). Again, the size of this

    term can also be expected to depend appreciably on factors

    such as solvent polarity. In particular, apolar solvents may lead

    to precipitation of [LB·H]+[LA·H]

    – salts; in these cases the ion

    pairing term becomes effectively very large, and may provide

    the main thermodynamic driving force for H2 activation.

    In general, of the five terms outlined above, only two are

    expected to vary very significantly upon variation of the FLP:

    H+ attachment and H

    – attachment. Thus, the thermodynamic

    ability of any FLP to activate H2 will critically depend on the

    magnitude of these terms, and hence with the combined

    ‘strength’ of the LB and LA; a conclusion that has been found

    to compare very well with experimental results. In particular, it

    can be seen that when designing FLP hydrogenation catalysts,

    the key measures by which the ‘strength’ of the LB and LA

    should be judged are their proton affinity (PA) and hydride ion

    affinity (HA), respectively. Experimentally-determined proxies

    for PA have been extensively tabulated in the form of pKa

    values.17

    Importantly, these are often available for a variety of

    different solvents (solvation having a significant effect not just

    on PA and pKa but also HA; in general, more polar solvents are

    expected to render FLP activation of H2 by neutral LA/LB

    combinations more favourable, by stabilising the ionic

    products relative to the neutral reactants). Experimental

    values for HA are unfortunately far less abundant and so

    alternative measures of LA strength are often used as

    alternatives to aid FLP design (for example the commonly-

    employed Gutmann-Beckett method, which uses changes in 31

    P chemical shift to probe the strength of binding between

    LAs and Et3PO).18

    Nevertheless, the relationships between HA

    and other measures of abstract ‘Lewis acidity’ or

    ‘electrophilicity’ are not always trivial and may show very

    different sensitivity to other relevant parameters such as steric

    bulk (other chemical probes will have different steric profiles

    than H–).

    19 As such, these values are often most reliable (and

    so most useful) as guidelines when comparing structurally-

    similar LAs that show variation in positions distant from the

    acidic centre, rather than for comparison of more diverse LAs

    from different ‘families’. It is also often useful to consider such

    values in conjunction with calculated estimates of HA;

    fortunately, values for a fairly diverse collection of FLP-

    relevant main-group LAs were recently reported by Heiden

    and Latham.20

    Taken together, PA and HA values afford an invaluable

    predictive tool for the design of FLP hydrogenation catalysts.

    By comparing the values for prospective FLPs with those of

    systems already reported in the literature, it is possible to

    anticipate their likely degree of reactivity towards H2. If the

    combined PA and HA are very low, then H2 activation will be

    highly disfavoured, and successful hydrogenation catalysis is

    likely to be infeasible. Conversely, for effective catalysis the

    combined PA and HA of the FLP should also not be excessively

    high, as this will lead to a very stable and hence unreactive

    [LB·H]+[LA·H]

    – H2 cleavage product. In such cases either H

    + or

    H– transfer (or both) to the substrate will be unfavourable, and

    turnover will again be limited. If the reaction mechanism

    involves LA activation of the substrate (Fig. 4b) this also

    requires that H2 activation be sufficiently reversible to ensure

  • ARTICLE Journal Name

    6 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

    Please do not adjust margins

    Please do not adjust margins

    the presence of some ‘free’ LA (assuming 1:1 LA:LB

    stoichiometry). It should be noted that a number of ‘weak’

    FLPs have actually been observed to effect successful catalytic

    hydrogenation despite not activating H2 to a sufficient extent

    for it to be observed by standard NMR spectroscopic

    techniques. In these cases the ability of a FLP to effect

    transient, reversible H2 cleavage can often be demonstrated by

    admitting HD gas or a mixture of H2 and D2, and observing

    isotopic scrambling to form the statistical 2:1:1 mixture of HD,

    H2 and D2 (Fig. 7). Thus, B(C6F5)3 has been found to effect

    successful catalytic hydrogenation in combination with LBs as

    weak as simple ethers (aqueous pKaH < 0).21

    In contrast, FLPs

    consisting of the same LA and very powerful LBs such as N-

    heterocyclic carbenes (NHCs; aqueous pKaH ~ 20-25) have not

    yet found use in hydrogenation catalysis, even though they

    readily activate H2.22

    In this context, it is noteworthy that the

    pKa of H2 itself has been experimentally estimated to be

    approximately 35 in THF.23

    In principle, the LA-free

    deprotonation of H2 could be thought of as a conceptual

    limiting case for FLP H2 activation, where a very large PA term

    is necessary in order to compensate for negligible HA.

    In addition to combined PA and HA, it is important that the

    PA affinity alone (and, in principle, HA alone) is tailored to

    those of the substrate and product (analogous to the

    electronic fine-tuning typically required of TM catalysts). As

    shown in Fig. 4a, the standard mechanism for FLP-catalysed

    hydrogenation requires that the [LB·H]+ intermediate be a

    sufficiently strong Brønsted acid to protonate the substrate (or

    else activate it appreciably through hydrogen-bonding), which

    places an upper limit on the strength of the LB that can be

    used. These principles were elegantly illustrated by Paradies

    and co-workers during the development of phosphine/B(C6F5)3

    FLPs for alkene hydrogenation, where it was found that sub-

    optimal rates were obtained when using phosphine LBs that

    were either too weak (where disfavourable H2 activation is

    rate-limiting) or too strong (where substrate protonation to

    form an intermediate carbocation becomes rate-limiting

    instead).24

    Notably, different optimum LB strengths were

    found when the basicity of the substrate was changed,

    highlighting the need to tailor LB strength to the specific

    substrate under investigation.

    The basicity of the substrate is of even more importance in

    LA-only catalyst systems, where it is directly involved in H2

    activation (Fig. 4c). Typically, this reaction pathway is feasible

    for more strongly basic substrates, while less basic analogues

    benefit strongly from addition of a stronger auxiliary LB which

    can speed H2 activation and act as an intermediate ‘proton

    shuttle’.

    In some cases it may also be important to consider the

    basicity of the intended reaction product. In a particularly

    extreme example, Stephan et al. reported that while B(C6F5)3 is

    capable of mediating the stoichiometric hydrogenation of

    simple anilines to cyclohexylamines, catalytic turnover is

    prohibited due to the high basicity of the product (amine pKaH

    ~ 10 in H2O), which prevents protonation of the much less

    basic aromatic substrate (aniline pKaH ~ 5 in H2O; note that

    while this value relates to protonation at nitrogen, the initial

    site of protonation required for reduction is actually at carbon,

    which will be less basic). In effect, the Brønsted acidity is

    ‘levelled’ to the weak cyclohexylammonium species (Fig. 8a).25

    Conversely, Paradies et al. have described a degree of

    autocatalysis in certain borane-catalysed imine

    hydrogenations.26

    This is attributed to the increased basicity of

    the product amines, which means that rate-limiting H2

    activation becomes more favourable as the reaction proceeds

    (Fig. 8b; both of these examples involve LA-only catalysis).

    Balancing steric bulk

    Perhaps the most obvious variables in the design of FLP

    catalysts are the steric bulk of the acidic and basic centres. The

    fundamental FLP concept clearly requires that the LA and LB

    possess sufficient combined bulk to prevent formation of a

  • Journal Name ARTICLE

    This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 7

    Please do not adjust margins

    Please do not adjust margins

    strong classical adduct. If these functional groups are too small

    then mutual quenching will eliminate the ability of the FLP to

    engage in the H—H cleavage step that is crucial to catalysis

    (even if PA/HA and other factors would otherwise be

    favourable, the unfavourable separation term in Fig. 6 will

    become insurmountably large). Steric bulk, particularly around

    the LA, is also a key factor in determining substrate scope and

    functional group tolerance for FLP hydrogenation catalysts.

    Soós and co-workers have emphasised the value of the ‘size-

    exclusion principle’ in FLP design, where the use of especially

    bulky LAs can allow for the effective hydrogenation of less

    bulky imine substrates, for example, by limiting unproductive

    adduct formation between the LA and product amines.27

    Use

    of a bulkier LA also means an FLP can be formed using a less

    bulky LB (or vice versa).

    Nevertheless, is it not simply the case that using bulkier

    components will lead to a more active FLP catalyst. Extremely

    bulky LAs will lead to similarly bulky [LA·H]– reductants, whose

    size can lead to low kinetic reactivity, particularly if the hydride

    needs to be transferred to a relatively bulky substrate. In

    particularly extreme cases steric bulk may even inhibit H2

    activation. For example, while investigating the very hindered

    LA MesB(C6F5)2, Soós et al. found that H2 activation was much

    slower when using LBs that were also very bulky (e.g. 2,2,6,6-

    tetramethylpiperidine), compared to when less hindered LBs

    of comparable pKa were employed (e.g. quinuclidine).27

    As a

    result, and perhaps counterintuitively, it can sometimes be

    productive to pursue the use of less bulky LAs, even if this

    appears to lead to the formation of a classical LA←LB adduct.

    Provided that any such dative interaction is reversible,

    transient cleavage can generate the active FLP in situ. In

    particular, Lewis pairs that appear to form a strong adduct at

    room temperature may dissociate significantly at elevated

    temperatures: this has come to be known as ‘thermally-

    induced frustration’.28

    While the need to separate the LA and

    LB provides an additional energetic barrier that must be

    overcome prior to H2 activation (the LA←LB separation term in

    Fig. 6 in no longer negligible), this is potentially outweighed by

    the kinetic advantages of generating a less bulky and more

    reactive [LA·H]– reductant (Fig. 9).

    Reversible adduct formation can also have significant

    implications for the stability of the catalytic Lewis pair. In

    particular, the decomposition of highly reactive FLPs may be

    suppressed by allowing them to form a reversible adduct. For

    example, while investigating the use of NHC LBs in FLP

    chemistry, Tamm et al. found that, while an NHC/B(C6F5)3 FLP

    undergoes rapid rearrangement to form an unreactive

    ‘abnormal NHC’ adduct over the course of a few hours at room

    temperature, the equivalent less hindered NHC/B(FXyl)3 pair

    [FXyl = 3,5-bis(trifluoromethyl)phenyl)] forms a reversible

    normal adduct that is stable up to significantly elevated

    temperatures, while still retaining its FLP-type reactivity

    towards H2 (Fig. 10a).22

    Taking the concept of thermally-induced frustration to its

    extreme, it can actually be possible for a classical adduct to

    display FLP-like reactivity without ever achieving complete

    separation of the Lewis pair. Ashley et al. have previously

    described how the adduct [iPr3Si←PtBu3]+[B(C6F5)4]

    –, which

    Fig. 10 Reversible adduct formation can stabilise FLPs against

    decomposition, for example in the systems shown in (a). FLP-like

    reactivity can also sometimes be observed by Lewis pairs without

    even transient separation (b). This can be compared with direct

    addition of H2 across certain polar chemical bonds (c; this is

    another strategy that has recently been used to achieve TM-free

    catalytic hydrogenation).

    Fig. 9 A qualitative summary of how catalytic activity can typically

    vary as a function of FLP steric bulk, assuming combined HA/PA

    and other parameters are suitable for catalysis. The ideal catalyst

    must be neither too large nor too small, but the ideal mid-point

    will be substrate-dependent.

  • ARTICLE Journal Name

    8 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

    Please do not adjust margins

    Please do not adjust margins

    does not dissociate to form PtBu3 + [iPr3Si]+[B(C6F5)4]

    – even at

    high temperature, is nevertheless capable of activating H2 in

    an FLP-like manner.29

    This is attributed to transient

    lengthening and weakening of the Si←P interaction, which

    generates a structure analogous to an FLP encounter complex

    that is capable of inserting H2. From here, it is possible to see

    an analogy between FLP H2 activation and the direct

    hydrogenolysis of certain weak or polar chemical bonds, which

    has also recently begun to be exploited to develop mild

    protocols for TM-free catalytic hydrogenation of alkenes (Fig.

    10b,c).30

    Choosing between intramolecular, intermolecular and ‘LA-only’

    FLPs

    As noted previously, tethering the LA and LB together to form

    an intramolecular FLP transforms H2 activation from a formally

    termolecular into a bimolecular reaction step, which is

    expected to have a much lower entropic barrier. As such,

    optimised intramolecular FLPs might intuitively be expected to

    be superior catalysts than their intermolecular counterparts

    (cf. TM catalysts, where the acidic and basic functionalities are

    by necessity also localised on a single molecule). Indeed, many

    investigations of intramolecular FLPs have appeared to support

    this conclusion. For example, in 2008 Erker et al. reported one

    of the first such systems: an ethylene-linked P/B system that

    showed far greater activity as an imine hydrogenation catalyst

    than previous intermolecular P/B FLPs.31

    Nevertheless, it is important to acknowledge that the

    development of intramolecular FLP catalysts suffers from

    some significant drawbacks. In particular, the activity of such

    systems is typically highly sensitive to the nature of the linker

    used to connect the LA and LB, in a manner that is not easily

    predictable. As such Erker et al., expanding on their earlier

    work, have shown that P/B FLPs with a variety of simple alkyl

    linkers show dramatically different reactivities towards H2.32

    For example, in the series of oligo(methylene) linked FLPs

    Mes2B(CH2)nP(C6F5)2, the members with n = 2 and n = 4 are

    both active hydrogenation catalysts, while the intermediate

    member with n = 3 is unreactive towards H2. Similarly, Aldridge

    and co-workers have reported that while a dimethylxanthene-

    linked P/B FLP readily activates H2 under mild conditions (room

    temperature, 1 bar H2), no such activation is observed using an

    otherwise identical dibenzofuran-linked system (Fig. 11).33

    Such differences in reactivity can be attributed to

    thermodynamic changes that arise when using different linkers

    (for example if the resulting ‘H+’ and ‘H

    –’ moieties are held too

    far apart for effective ionic stabilisation) or, in other cases, to

    kinetic factors that relate to the ease with which different

    intramolecular FLPs are able to preorganise themselves into a

    conformation suitable for H2 activation. In extreme cases, if no

    such conformation is energetically accessible, then formally

    intramolecular FLPs may in fact only activate H2 in an

    intermolecular fashion, as is the case for the original, rigid

    systems reported by Stephan et al. (see Fig. 2a).34

    Further complicating the design of intramolecular FLPs, the

    choice of linker will also directly impact both the steric and

    electronic properties of the attached LA and LB centres, which

    can lead to difficulties in their rational fine-tuning.

    Intramolecular FLPs also typically require longer, more

    demanding synthetic routes for their preparation than

    unlinked LAs and LBs (of which many of the most commonly

    used are commercially available). Thus, the expectation that

    intramolecular FLPs might ultimately provide superior catalysis

    must be balanced against the much greater ease and speed

    with which intermolecular catalysts can be developed (note

    that translating a successful intermolecular catalyst into an

    intramolecular analogue is also not necessarily trivial, for

    similar reasons). In particular, intermolecular FLPs lend

    themselves very well to screening efforts that can rapidly

    identify promising catalytic leads (especially valuable when the

    optimum catalyst is highly substrate-dependent). By contrast,

    even a rather modest screen of five acidic and five basic

    moieties would require the synthesis of 50 separate

    intramolecular FLPs, even if only two possible linkers were

    investigated. Mechanistic investigations are also often simpler

    using intermolecular systems, for similar reasons, as it is

    easier to synthesise possible intermediates, or to exclude

    either the acidic or basic component (cf. Fig 5; for an

    intramolecular system these investigations would require the

    further synthesis of separate monofunctional model

    compounds).

    The advantages enjoyed by intermolecular FLPs are even

    more apparent in the ‘LA-only’ catalytic systems discussed

    previously, where only one reaction component needs to be

    varied and the overall reaction mixture is significantly

    simplified. For example, the groups of Stephan and Crudden

    have shown how simple screening studies can be used to

    rapidly optimise the structure of borenium LAs for imine

    hydrogenation catalysis.35,36

    Again, however, this advantage

    must be weighed against the expectation that, with fewer

    variables available for optimisation, the best LA-only system

    may be less effective than the best inter- or intramolecular

    equivalent. As an example, it has been reported that, while

    hydrogenation of weakly-basic imines such as PhCH=NSO2Ph

    can proceed using B(C6F5)3 as the sole catalyst, greatly

  • Journal Name ARTICLE

    This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 9

    Please do not adjust margins

    Please do not adjust margins

    improved rates can be achieved upon addition of P(Mes)3 as a

    co-catalytic LB.37

    Understanding and controlling inhibition and decomposition

    pathways

    Just as important as designing an FLP that will be able to

    perform all the component steps of a catalytic cycle (H2

    activation; H+ and H

    – transfer) is ensuring that the potential

    catalyst can avoid any irreversible inhibition or decomposition

    steps that might hinder catalysis. Several general

    considerations have already been mentioned in earlier

    sections with respect to catalyst inhibition. For example, steric

    tuning can be used to avoid adduct formation between the

    LA/LB and any other basic or acidic functional groups present,

    while reversible adduct formation can be exploited to help

    stabilise overly reactive FLPs.

    A large number of chemical bonds aside from H—H are

    known to be susceptible to cleavage by FLPs, which can lead to

    quenching of the LA and LB. As such, if any of these

    functionalities are present in the substrate and/or product

    then efforts should be made to minimise the reactivity of the

    FLP towards them. In practice, given the overwhelming

    reliance of FLP hydrogenation catalysis on boron-based LAs

    and early p-block LBs, reducing reactivity towards other

    functional groups without a concomitant reduction in

    reactivity towards H2 is not always easy, and may require the

    use of less typical FLP components. An example of this will be

    discussed in the case study at the end of this review.

    Inevitably, relevant decomposition routes will be highly

    dependent on the precise system under study. Nevertheless,

    because most of the FLP hydrogenation catalysts reported to

    date are based on bulky, electrophilic fluoroaryl-substituted

    boranes it is worth specifically considering the decomposition

    of these compounds, which typically involves loss of an aryl

    group to an electrophile (Fig. 12a). A specific example is the

    decomposition of the ubiquitous B(C6F5)3 in the presence of

    simple alcohols (or H2O) which is believed to occur via initial

    coordination to form ROH→B(C6F5)3 adducts in which the O—H

    has been dramatically acidified, prompting intramolecular

    protodeborylation (Fig. 12b; note that the decomposition

    products ROBAr2 have much lower HAs, so are considerably

    less likely to activate H2 than the initial BAr3).38,39

    In general,

    aryl group transfer of this type should be particularly facile

    from anionic, 4-coordinate borate intermediates. Any catalytic

    protocol that involves a build-up of such intermediates in the

    presence of a suitable electrophile (which may simply be more

    of the initial borane LA, leading to sequestration of a second

    equivalent of borane in the form of a BAr4– anion) is therefore

    likely to be particularly susceptible to decomposition,

    particularly if the reaction is required to run at significantly

    elevated temperatures.

    Case study: C=O hydrogenation and moisture tolerance

    Over the past several years we have focused on the

    development of some of the first FLP systems capable of

    catalysing the hydrogenation of aldehydes and ketones, and of

    moisture-tolerant FLP hydrogenation catalysis. In this section

    we will discuss the rational development of these systems in

    order to illustrate the use of the principles that have been

    outlined so far (and which are summarised in the learning

    points listed at the start of this review).

    The first goal of our work was to develop a protocol for

    FLP-catalysed hydrogenation of aldehydes and ketones to

    alcohols, which had not previously been reported. Based on

    the principles outlined above it was possible to identify the

    following key factors:

    The standard mechanism for FLP-catalysed hydrogenation

    requires protonation of the substrate prior to H– transfer.

    Because organic carbonyls are very poor bases (aqueous pKa

    < 0), this suggests the LB employed in the FLP must also be

    very weak. This in turn means a LA with fairly high HA will

    be necessary so that H2 activation is feasible.

    Unlike isoelectronic imines and amines, organic carbonyls

    and alcohols lack steric bulk around their Lewis basic

    oxygen atoms, which suggests that relatively bulky LAs

    might be desirable to avoid inhibitory product→LA and

    substrate→LA adduct formation. Conversely, however,

    these neutral oxygen bases are much weaker donors than

    their nitrogen-based counterparts, which could mitigate this

    issue, and overly bulky [LA·H]– anions might lead to slow

    hydride transfer. It therefore seemed sensible to investigate

    LAs with a range of steric profiles.

    The use of ‘LA-only’ systems for stoichiometric carbonyl

    hydrogenation had been reported previously, using B(C6F5)3

    as the catalyst.38

    These investigations had confirmed that

    direct H2 activation using the substrate as LB is feasible;

    however, decomposition of the borane under these highly

    acidic conditions meant that turnover could not be

  • ARTICLE Journal Name

    10 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

    Please do not adjust margins

    Please do not adjust margins

    observed. It seemed likely that a stronger auxiliary LB might

    lead both to improved H2 activation kinetics, and to more

    stable Brønsted acidic intermediates. It was also speculated

    that further improvements in stability might be observed

    for a system with ‘thermally induced frustration’. Because

    the desired reactivity lacks precedent, the initial goal was to

    obtain a system that is catalytically competent, prior to full

    optimisation. It therefore seemed sensible to investigate

    intermolecular rather than intramolecular systems.

    Though some FLPs are known to activate C=O bonds, it was

    judged that the most significant inhibitory side-reaction was

    likely to be activation of the product O—H bonds, as

    previous borane-based FLPs had been reported to be very

    sensitive to inhibition by hydroxylic species, including

    alcohols.37

    While this had typically been attributed in

    general terms to the oxophilicity of boron, a closer analysis

    of the literature suggested that the more specific problem

    relates to irreversible deprotonation of the highly Brønsted

    acidic borane-alcohol adduct (aqueous pKa < 0; Fig. 13).39

    Developing boron-based FLP catalysts40

    Based on the above considerations, and given the well-

    established utility of highly electrophilic triarylborane LAs in

    FLP hydrogenation catalysis, it was decided to begin

    investigations by examining the boranes B(C6Cl5)n(C6F5)3-n (n =

    0-3) whose chemistry we had investigated previously, and with

    which we were therefore familiar.41

    These also satisfied the

    requirement of having large HAs and a range of different steric

    bulk. To begin with, preliminary mechanistic investigations

    were carried out in which pre-formed [Bu4N]+[HBAr3]

    – salts

    were reacted with a model substrate (acetone). These

    confirmed that direct reactions of [HBAr3]– with the substrate,

    either alone or in the presence of additional BAr3, were

    unlikely to be feasible mechanistic steps (cf. Fig. 5). Thus,

    catalytic hydrogenation would have to proceed via the

    standard mechanism (Fig. 4a), which involves protonation of

    the substrate. It was therefore decided to combine the initial

    borane LAs with simple ethers as LBs (aqueous pKaH < 0; Fig.

    14). The use of such weak bases should ensure that substrate

    activation through protonation or hydrogen-bonding is

    feasible, while also preventing irreversible deprotonation of

    any ROH→BAr3 adducts (cf. Fig. 13). Reversible coordination

    between the ethers and boranes might also reduce any

    tendency towards decomposition (e.g. via alcoholysis; cf. Fig.

    12b). Furthermore, combinations of this type had previously

    been shown to be effective in the hydrogenation of other

    weakly-basic substrates.41

    Finally, because these LBs are

    cheaply commercially available they are well-suited towards

    catalyst screening efforts, and because they can be used as

    solvents the entropic penalty faced by intermolecular FLPs

    when activating H2 can be minimised.

    The system chosen for initial investigation was the

    B(C6Cl5)(C6F5)2/THF Lewis pair; previously studies had shown

    that adduct formation in this system is highly reversible,

    leading to effective catalysis in the hydrogenation of other

    substrates.41

    As hoped, this system was also able to achieve

    successful catalysis in the hydrogenation of a model ketone

    (acetone), but turnover in the reaction was unexpectedly

    limited due to side-reactions of the product alcohol.

    Fortunately, subsequent screening (made simple by the

    decision to pursue an intermolecular system) was able to

    rapidly identify commercially-available B(C6F5)3 and 1,4-

    dioxane solvent as a superior protocol that avoids this issue.42

    Having confirmed that catalytic systems of this type are

    capable of tolerating simple alcohols, it was realised that

    similar systems might also be tolerant of another significant

    hydroxylic inhibitor of FLP catalysis: H2O (low air and moisture

    tolerance have typically been severe limitations of early FLP

    catalysis). This was quickly confirmed, although reactions were

    observed to be significantly slower than under strictly

    anhydrous conditions, which necessitated the use of increased

    H2 pressures.43

    Indeed, closer inspection of both the ‘wet’ and

    anhydrous reactions suggested that, despite their apparent

    tolerance, activation of O—H bonds remains a significant

    limiting factor in the rates of these reactions (as indicated, for

  • Journal Name ARTICLE

    This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 11

    Please do not adjust margins

    Please do not adjust margins

    example, by 11

    B NMR spectroscopy), even in the absence of

    any very strong bases. Note that increased temperatures could

    not be used to increase reaction rate as significant

    hydrolysis/alcoholysis of the borane was observed in these

    cases, consistent with the previously-anticipated

    decomposition route.

    The switch to Sn-based FLP catalysis44

    Development of the borane-based systems described above

    had emphasised the importance of O—H cleavage as a

    mechanism for inhibition of the FLP catalyst. Because boron is

    particularly oxophilic, even moderately strong LBs will cause

    this reaction to be irreversible when using borane LAs. This

    severely limits the scope of the ‘OH-tolerant’ reactions that

    had been developed, as neither the substrates nor products

    can contain any appreciably basic functional groups (Fig. 15).

    The use of very weak LBs also leads to the formation of very

    powerful [LB·H]+ Brønsted acids after H2 cleavage, which

    means that highly acid-sensitive functional groups also have to

    be avoided.

    In order to overcome these limitations and develop

    protocols that include and tolerate stronger Lewis bases the

    following further key factors were identified:

    To minimise the problem of unproductive RO—H activation,

    LAs should be incorporated that bind more weakly to –OR

    moieties. However, in order to maintain the ability of the

    FLP to activate H2, the HA of the LA must be preserved. In

    other words, a LA is needed that is relatively less oxophilic,

    but still hydridophilic.

    The presence of stronger LBs will prevent protonation of

    much more weakly basic substrates such as aldehydes and

    ketones. Hydrogenation of such compounds is therefore

    likely to require an atypical reaction mechanism, such as

    activation of the substrate by the LA rather than H+ (Fig.

    4b).

    Based on this analysis, subsequent work was directed towards

    investigating the use of previously-unexplored R3SnOTf LAs (R

    = alkyl; these are surrogates for [R3Sn]+ cations, which are

    valence isoelectronic with BAr3), in combination with typical

    nitrogen LBs. Inspection of the literature indicated that these

    acids should be significantly less oxophilic than boranes of

    similar HA (such as B(C6F5)3),20

    as indicated by the aqueous

    pKa values of the respective LA·OH2 adducts (< 0 for B(C6F5)3

    versus ~6 for [Bu3Sn]+; note that this is a specific example

    where it is important to consider a measure of LA strength

    other than just HA). In addition, R3SnOTf-catalysed addition of

    R3SnH (which is generated following H2 activation by R3SnOTf-

    based FLPs) to aldehydes and ketones is well established,

    which suggested that the necessary atypical reaction

    mechanism should be feasible (simple mechanistic

    investigations were again performed to confirm this).

    Gratifyingly, an optimised FLP consisting of iPr3SnOTf and

    2,4,6-collidine was found to be capable not only of activating

    H2, but also of both hydrogenating aldehydes and ketones and

    performing hydrogenation catalysis in the presence of

    moisture, despite the presence of fairly strong Brønsted

    bases.

    Conclusions and future outlook

    Despite the enormous advances that have been made in field

    of FLP catalysis in the decade since it was established, there

    remains significant scope for further development. It is our

    hope that the principles outlined in this review may provide a

    useful framework for this ongoing work.

    While the systems reported to date have provided a

    dramatic proof-of-principle for TM-free catalytic

    hydrogenation, it would be hard to argue that any of these

    reactions yet constitutes a truly attractive synthetic tool. If

    FLPs are eventually to take their place alongside TM complexes

    as practical hydrogenation catalysts then future work must

    increasingly be focused on ensuring broad functional group

    tolerance, ‘open bench’ stability and well-defined

    chemoselectivity, alongside optimised reactivity. In this vein, it

    has been encouraging to see recent reports focusing on the

    development of catalytic protocols that can achieve high

    enantioselectivity,45

    that use very low catalyst loadings,35

    or

    that catalyse reactions for which there is no existing TM-

    catalysed alternative.46

    Based on our own experiences with Sn-based systems, we

    would suggest that achieving the above goals may require a

    willingness to investigate a broader range of LAs and LBs than

    have currently been investigated.47

    In particular, s-block, late

    p-block, and cheap d-block LAs all remain relatively unexplored

    for applications in catalytic FLP chemistry. Detailed

    experimental investigations into the kinetics and mechanisms

    of most FLP-catalysed hydrogenation reactions are also still

    needed, to provide the theoretical basis for rational future

    development. Finally, it may be possible that FLP H2 activation

    could find application in areas other than chemical

    hydrogenation catalysis. Plausible examples include reversible

  • ARTICLE Journal Name

    12 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

    Please do not adjust margins

    Please do not adjust margins

    hydrogen storage using very low molecular-weight FLPs, and

    electrocatalytic H2 oxidation reactions for use in hydrogen fuel

    cell applications.48

    Acknowledgements

    We would like to thank GreenCatEng, Eli Lilly (Pharmacat

    consortium) and the EPSRC for providing funding (D. J. S.) and

    the Royal Society for a University Research Fellowship (A. E.

    A.).

    Notes and references

    1 For an excellent summary of this area, see P. P. Power, Nature, 2010, 464, 171. See also cited and citing references.

    2 G. D. Frey, V. Lavallo, B. Donnadieu, W. W. Schoeller and G. Bertrand, Science, 2007, 316, 439.

    3 G. C. Welch, R. R. S. Juan, J. D. Masuda and D. W. Stephan, Science, 2006, 314, 1124.

    4 G. C. Welch and D. W. Stephan, J. Am. Chem. Soc., 2007, 129, 1880.

    5 G. C. Welch, L. Cabrera, P. A. Chase, E. Hollink, J. D. Masuda, P. Wei and D. W. Stephan, Dalton Trans., 2007, 3407.

    6 The field of FLP chemistry has seen a number of broad summaries in recent years. See, for example, D. W. Stephan, Science, 2016, 354, 1248, and references therein.

    7 An alternative, electric-field-based model was also proposed, but has largely fallen out of favour and is not regularly invoked. See, for example, T. A. Rokob, I. Bakó, A. Stirling, A. Hamza and I. Pápai, J. Am. Chem. Soc., 2013, 135, 4425.

    8 P. A. Chase and D. W. Stephan, Angew. Chem. Int. Ed., 2008, 47, 7433.

    9 See W. E. Piers, A. J. V. Marwitz and L. G. Mercier, Inorg. Chem., 2011, 50, 12252, and references therein.

    10 J. R. Khusnutdinova and D. Milstein, Angew. Chem. Int. Ed., 2015, 54, 12236.

    11 K. K. Ghuman, L. B. Hoch, T. E. Wood, C. Mims, C. V. Singh and G. A. Ozin, ACS Catal., 2016, 6, 5764.

    12 See A. Berkessel, T. J. S. Schubert and T. N. Müller, J. Am. Chem. Soc., 2002, 124, 8693, and references therein.

    13 For a recent review, see R. H. Morris, Chem. Rev., 2016, 116, 8588.

    14 T. A. Rokob, A. Hamza, A. Stirling, T. Soós and I. Pápai, Angew. Chem. Int. Ed., 2008, 47, 2435.

    15 L. Rocchigiani, G. Ciancaleoni, C. Zuccaccia and A. Macchioni, J. Am. Chem. Soc., 2014, 136, 112.

    16 T. A. Rokob, A. Hamza and I. Pápai, J. Am. Chem. Soc., 2009, 131, 10701.

    17 See, for example: F. G. Bordwell, Acc. Chem. Res., 1988, 21, 456 and cited and citing references. Note that where pKa values are given in this review, they have typically been taken from the related cited work (or references therein).

    18 M. A. Beckett, G. C. Strickland, J. R. Holland and K. Sukumar Varma, Polymer, 2996, 37, 4629.

    19 See, for example: A. E. Ashley, T. J. Herrington, G. G. Wildgoose, H. Zaher, A. L. Thompson, N. H. Rees, T. Krämer and D. O’Hare, J. Am. Chem. Soc., 2011, 133, 14727; or É. Dorkó, B. Kótai, T. Földes, Á. Gyömöre, I. Pápai and T. Soós, J. Organomet. Chem., 2017, doi: 10.1016/j.jorganchem.2017.04.031.

    20 Z. M. Heiden and A. P. Latham, Organometallics, 2015, 34, 1818.

    21 L. J. Hounjet, C. Bannwarth, C. N. Garon, C. B. Caputo, S. Grimme and D. W. Stephan, Angew. Chem. Int. Ed., 2013, 52, 7492.

    22 See E. L. Kolychev, T. Bannenberg, M. Freytag, C. G. Daniliuc, P. G. Jones and M. Tamm, Chem. Eur. J., 2012, 18, 16938, and references therein.

    23 See E. Buncel and B. Menon, J. Am. Chem. Soc., 1977, 99, 4457, and references therein.

    24 See L. Greb, S. Tussing, B. Schirmer, P. Ona-Burgos, K. Kaupmees, M. Lokov, I. Leito, S. Grimme and J. Paradies, Chem. Sci., 2013, 4, 2788, and references therein.

    25 T. Mahdi, Z. M. Heiden, S. Grimme and D. W. Stephan, J. Am. Chem. Soc., 2012, 134, 4088.

    26 See S. Tussing, K. Kaupmees and J. Paradies, Chem. Eur. J., 2016, 22, 7422, and references therein.

    27 See T. Soós, Pure Appl. Chem., 2011, 83, 667, and references therein.

    28 T. A. Rokob, A. Hamza, A. Stirling and I. Pápai, J. Am. Chem. Soc., 2009, 131, 2029.

    29 T. J. Herrington, B. J. Ward, L. R. Doyle, J. McDermott, A. J. P. White, P. A. Hunt and A. E. Ashley, Chem. Commun., 2014, 50, 12753.

    30 See, for example, Y. Wang, W. Chen, Z. Lu, Z. H. Li and H. Wang, Angew. Chem. Int. Ed., 2013, 52, 7496.

    31 P. Spies, S. Schwendemann, S. Lange, G. Kehr, R. Fröhlich and G. Erker, Angew. Chem. Int. Ed., 2008, 47, 7543.

    32 See T. Özgün, K.-Y. Ye, C. G. Daniliuc, B. Wibbeling, L. Liu, S. Grimme, G. Kehr and G, Erker, Chem. Eur. J., 2016, 22, 5988, and references therein.

    33 Z. Mo, E. L. Kolychev, A. Rit, J. Campos, H. Niu and S. Aldridge, J. Am. Chem. Soc., 2015, 137, 12227.

    34 Y. Guo and S. Li, Inorg. Chem., 2008, 47, 6212. 35 J. M. Farrell, R. T. Posaratnanathan and D. W. Stephan,

    Chem. Sci., 2015, 6, 2010. 36 P. Eisenberger, B. P. Bestvater, E. C. Keske and C. M.

    Crudden, Angew. Chem. Int. Ed., 2015, 54, 2467. 37 See, for example, D. W. Stephan, S. Greenberg, T. W.

    Graham, P. Chase, J. J. Hastie, S. J. Geier, J. M. Farrell, C. C. Brown, Z. M. Heiden, G. C. Welch and M. Ullrich, Inorg. Chem., 2011, 50, 12338, and references therein.

    38 See L. E. Longobardi, C. Tang and D. W. Stephan, Dalton Trans., 2014, 43, 15723, and references therein.

    39 C. Bergquist, B. M. Bridgewater, C. J. Harlan, J. R. Norton, R. A. Friesner and . Parkin, J. Am. Chem. Soc., 2000, 122, 10581.

    40 In parallel with our own work described in this case study, very similar boron-based systems that address the same issues were independently developed by the groups of Stephan and Soós, and published near-simultaneously. See T. Mahdi and D. W. Stephan, J. Am. Chem. Soc., 2014, 136, 15809; and Á. Gyömöre, M. Bakos, T. Földes, I. Pápai, A. Domján and T. Soós, ACS Catal., 2015, 5, 5366.

    41 See D. J. Scott, M. J. Fuchter and A. E. Ashley, Angew. Chem. Int. Ed., 2014, 53, 10218, and references therein.

    42 D. J. Scott, M. J. Fuchter and A. E. Ashley, J. Am. Chem. Soc., 2014, 136, 15813.

    43 D. J. Scott, T. R. Simmons, E. J. Lawrence, G. G. Wildgoose, M. J. Fuchter and A. E. Ashley, ACS Catal., 2015, 5, 5540.

    44 D. J. Scott, N. A. Phillips, J. S. Sapsford, A. C. Deacy, M. J. Fuchter and A. E. Ashley, Angew. Chem. Int. Ed., 2016, 55, 14738.

    45 For a recent review of enantioselective FLP-catalysed hydrogenation, see J. Paradies, Chiral Borane-Based Lewis Acids for Metal Free Hydrogenations, Top. Curr. Chem., Springer GmbH, Berlin, 2017, pp 1-24.

    46 S. Wei and H. Du, J. Am. Chem. Soc., 2014, 136, 12261. 47 For a review of less standard main-group LAs in FLP

    chemistry, see S. A. Weicker and D. W. Stephan, Bull. Chem. Soc. Jpn., 2015, 88, 1003.

    48 See, for example, E. J. Lawrence, V. S. Oganesyan, D. L. Hughes, A. E. Ashley and G. G. Wildgoose, J. Am. Chem. Soc., 2014, 136, 6031.