29
University of Groningen Carbon-nitrogen bond formation via catalytic alcohol activation Yan, Tao IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 2017 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Yan, T. (2017). Carbon-nitrogen bond formation via catalytic alcohol activation. University of Groningen. Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons). Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 03-06-2021

University of Groningen Carbon-nitrogen bond formation via ......Chapter 2 24 Catalytic N-alkylation with alcohols Preliminary experiments using 4-methoxyaniline (1a) and a simple

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • University of Groningen

    Carbon-nitrogen bond formation via catalytic alcohol activationYan, Tao

    IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

    Document VersionPublisher's PDF, also known as Version of record

    Publication date:2017

    Link to publication in University of Groningen/UMCG research database

    Citation for published version (APA):Yan, T. (2017). Carbon-nitrogen bond formation via catalytic alcohol activation. University of Groningen.

    CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

    Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

    Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

    Download date: 03-06-2021

    https://research.rug.nl/en/publications/carbonnitrogen-bond-formation-via-catalytic-alcohol-activation(ae62681f-b95e-49bd-a23d-5cb5cecfb894).html

  • 21

    Chapter 2

    Iron catalyzed direct alkylation of amines with alcohols

    The selective conversion of carbon-oxygen bonds into carbon-nitrogen bonds to

    form amines is one of the most important chemical transformations for the

    production of bulk and fine chemicals and pharma intermediates. An attractive

    atom economic way of carrying out such C-N bond formations is the direct N-

    alkylation of simple amines with alcohols through the so-called borrowing

    hydrogen strategy. Recently, transition metal complexes based on precious noble

    metals have emerged as suitable catalysts for this transformation; however, the

    crucial change towards highly selective methodologies, which use abundant,

    inexpensive and environmentally friendly metals, in particular iron, has not yet

    been accomplished. In this chapter, the homogeneous, iron-catalyzed, direct

    alkylation of amines with alcohols is described. The scope of this new methodology

    includes the selective monoalkylation of anilines and benzyl amines with a wide

    range of alcohols as well as the use of diols in the formation of five-, six- and

    seven- membered nitrogen heterocycles, which are privileged structures in

    numerous pharmaceuticals.

    Part of this chapter was published:

    T. Yan, B. L. Feringa, K. Barta, Nat. Commun., 2014, 5, 5602.

  • Chapter 2

    22

    Introduction

    Amines are among the most valuable classes of compounds in chemistry,

    omnipresent in natural products, in particular alkaloids[2a], and widely used as

    pharmaceuticals, agrochemicals, lubricants and surfactants[1,2,3]. Therefore the

    development of efficient catalytic methodologies for C-N bond formation is a

    paramount goal in organic chemistry.

    The choice of an alcohol[4,5] as substrate for direct C-N bond formation is highly

    desirable in order to produce secondary and tertiary amines and N-heterocyclic

    compounds (Figure 1).

    Figure 1: Catalytic, direct N-alkylation of amines with alcohols. a Conventional conversion

    of alcohol into amine via installing a leaving group before nucleophilic substitution with an

    amine donor or oxidizing alcohols to carbonyl compounds followed by reductive amination.

    b The direct C–N bond formation via coupling of primary alcohols and amines forms

    secondary amine products R1–NH–R2. R1–OH is a short- or long-chain aliphatic alcohol and

    may contain aromatic functionality. R2–NH2 is an aromatic or aliphatic amine. c Using diols

    of various chain lengths the products are 5- (n=1), 6- (n=2) or 7- (n=3) membered N-

    heterocycles.

    Alcohols are readily available through a variety of industrial processes and are

    highly relevant starting materials in view of recent developments in the field of

    renewables as they can be obtained via fermentation or catalytic conversion of

    lignocellulosic biomass.[6,7] Conventional non-catalytic transformations of an amine

    with an alcohol take place via installing a suitable leaving group instead of the

    alcohol functionality followed by nucleophilic substitution, or oxidizing alcohols to

    carbonyl compounds followed by reductive amination; these multistep pathways

    suffer from low atom economy[8] or limited selectivity and the production of

    stoichiometric amounts of waste (Figure 1a).[9]

    A privileged catalytic methodology for the direct coupling of alcohols with amines

    is based on the so-called borrowing hydrogen strategy (Figure 1b, 1c and Figure

    2). During the catalytic cycle an alcohol is dehydrogenated to the corresponding

    carbonyl compound, which reacts with the amine to form an imine. The imine is in

  • Alkylation of amines with alcohols

    23

    situ reduced to the alkylated amine and the metal complex facilitates the required

    hydrogen shuttling. In fact the hydrogen delivered by the alcohol is temporarily

    stored in the metal complex. Key features are that the process is hydrogen neutral,

    no other reagents are needed and the only stoichiometric by-product is water.

    A number of transition metal complexes have proven effective in this catalytic C-

    N bond formation, in particular those based on ruthenium[10-12] and iridium[13].

    These and related methodologies have been extensively reviewed[1,3,4,14-16].

    Despite recent progress, the key challenge is the development of catalysts that

    rely on the use of widely abundant, inexpensive metals[3,17]. Iron is considered to

    be the ultimate, sustainable alternative for ruthenium[18], however, no unequivocal

    reductive amination via a borrowing hydrogen mechanism has been reported[19].

    Direct N-alkylation of amines with alcohols is limited to iron-halogenides under

    rather harsh reaction conditions (160-200 °C), not proceeding via a hydrogen

    autotransfer pathway[20].

    This work shows that a well-defined homogenous Fe-based catalyst can be

    successfully used in this atom-economic process, with a broad substrate scope.

    These direct, Fe-catalyzed transformations are highly modular, and provide water,

    as the only stoichiometric byproduct. The products are valuable secondary and

    tertiary amines or heterocycles, which contain diverse moieties R1 (from the

    alcohol substrate) and R2 (from the amine substrate) (Figure 1).

    The presented direct waste-free alcohol to amine functional group interconversions

    are important toward the development of sustainable iron based catalysis and will

    enable the valorization of biomass-derived alcohols in environment-benign

    reaction media.

    Results and discussion

    We reasoned that direct C-N bond formation with an iron catalyst is possible

    provided by the catalytic complex which shows high activity both in alcohol

    dehydrogenation (Figure 2a, Step 1) and imine hydrogenation (Figure 2a, Step 3).

    The realization of this concept for direct alkylation of amines with alcohols using

    cyclopentadieneone iron tricarbonyl complex Cat 3, the precurse to form Knölker’s

    complex[21] Cat 3-H, is presented here (Figure 2). Iron cyclopentadienone complex

    Cat 3 (Figure 2b), has been recently employed in catalysis including hydrogenation

    of ketones[22], reductive amination[23], transfer hydrogenation of carbonyl

    compounds and imines[24a], Oppenauer-type oxidation of alcohols[24b] as well as

    cooperative dual catalysis[25]. Given its unique reactivity, we considered Cat 3 a

    promising candidate for the development of the desired iron-catalyzed hydrogen

    borrowing methodology. To achieve the direct amination of alcohols, the key

    challenge is to match alcohol dehydrogenation and imine hydrogenation steps by

    establishing conditions under which the formed Fe-H species (Cat 3-H) from the

    initial alcohol dehydrogenation step is able to reduce the imine at a sufficient rate,

    not requiring the use of dihydrogen.

  • Chapter 2

    24

    Catalytic N-alkylation with alcohols Preliminary experiments using 4-

    methoxyaniline (1a) and a simple aliphatic alcohol, 1-pentanol (2a), with 5 mol%

    pre-catalyst Cat 3 and 10 mol% Me3NO oxidant (to form active Cat 3-O) at 110 °C

    in toluene showed the formation of 4-methoxy-n-pentylaniline (3aa), albeit with

    only 30% selectivity at 48% substrate conversion (Table 1, entry 1). Although

    promising, the initial results using common organic solvents indicated low

    conversion of 1a or low selectivity towards 3aa and analysis of the reaction

    mixtures identified insufficient imine reduction as the key problem. We reasoned

    that most probably a weakly coordinating ethereal solvent is needed to stabilize

    the key iron intermediates Cat 3-O[27]. In addition, imine formation might be

    facilitated in solvents, which have limited miscibility with water. The major

    breakthrough came when the green solvent cyclopentyl methyl ether (CPME), one

    that uniquely combines such properties, was selected as reaction medium. CPME

    recently emerged as a low-toxicity, sustainable solvent alternative for

    tetrahydrofuran[28].

    Figure 2 Individual reaction steps in the iron-catalyzed N-alkylation of amines using iron

    cyclopentadienone complexes. a The overall transformation is the direct coupling of an

    alcohol (R1–OH) with an amine (R2–NH2) to form the product amine (R1–NH–R2). The

    sequence of reaction steps starts with the dehydrogenation of R1–OH to the corresponding

    aldehyde with iron complex Cat 3-O (Step 1). Thereby one ‘hydrogen equivalent’ is

    temporarily stored at the bifunctional iron complex Cat 3-O, which is converted to its

    reduced, hydride form Cat 3-H. In Step 2, the carbonyl intermediate reacts with amine

    R2–NH2 to form an imine intermediate and water. In Step 3, the hydrogen equivalent

    “borrowed” from alcohol R1–OH are used in the reduction of the imine intermediate to

    obtain the desired product R1–NH–R2. The reduction is accompanied by conversion of iron

    hydride Cat 3-H to Cat 3-O, thereby a vacant coordination site is regenerated and the

    catalytic cycle closed. b Reactivity of Fe cyclopentadienone complexes: Cat 3 is an air- and

    moisture-stable iron tricarbonyl precatalyst. The active Cat 3-O complex is in situ

    generated from Cat 3 by the addition of Me3NO to remove one CO. Cat 3-O is readily

    converted to Cat 3-H by reaction with an alcohol.

  • Alkylation of amines with alcohols

    25

    Table 1: Optimization of reaction conditions for N-alkylation of p-methoxyaniline

    (1a) with 1-pentanol (2a).

    Entry 2a

    [eq.]

    Sol. T

    [°C]

    Conv. 1a

    [%]a

    Sel. 3aa

    [%]a

    Sel. 4aa

    [%]a

    1 1 Toluene 110 48 30 10

    2 2 Toluene 110 71 50 20

    3 2 DCE 110 75 15 41

    4 2 CH3CN 110 10

  • Chapter 2

    26

    Table 2: Product formation profiles for products 3aa, 4aa and substrate 1a.

    Entry Time [h] Remain 1a

    [%]a

    Sel. 3aa

    [%]b

    Sel. 4aa

    [%]b

    1 1 92 11 6

    2 2 59 31 8

    3 3 48 40 8

    4 5 21 70 4

    5 7 1 92 1

    General reaction conditions: General Procedure A, 1a (0.5 mmol), 2a (1 mmol), 130 °C,

    CPME (2 ml). Reactions were set up in parallel and runs were stopped at given time. aConversion determined based on GC-FID using octadecane as internal standard. bSelectivity determined based on GC-FID and corresponding conversion.

    Next, an in situ NMR study (Figure 3) was conducted, using d-8 toluene at 100 °C

    that allowed the detection of all key reaction intermediates (as depicted in Figure

    2a), that is, 1-pentanol (2a), 1-pentanal, amine 3aa and the corresponding imine,

    in support of the borrowing hydrogen mechanism.

    Selective monoalkylation of anilines The general applicability of this method

    for the selective monoalkylation of substituted anilines was examined using 1-

    pentanol 2a and 17 anilines 1a–1q with diverse electron density and steric

    hindrance on the amino group. The isolated yields under optimized conditions are

    shown in Table 3.

  • Alkylation of amines with alcohols

    27

    Figure 3: In situ 1H NMR study of the N-alkylation of p-methoxyaniline (1a) with 1-

    pentanol (2a) in toluene-d8 at 100 °C. Full ppm range, showing 2 small absorptions at

    9.41 and 7.65 ppm, which can be attributed to HA (pentane-1-al)[26a] and HI (imine 4aa)[26b].

    Table 3: Selective monoalkylation of anilines with alcohols.

    General reaction conditions: General Procedure A, 0.5 mmol 1, 1 mmol 2, 2 ml CPME,

    130 °C, 18 h, isolated yields, unless otherwise specified. For details see Table 4 and Table

    5. a120 °C; bselectivity determined by GC-FID; c2 mmol 2 was used.

  • Chapter 2

    28

    Monoalkylation of various anilines with 1-pentanol The reactivity of

    substrates 1a–1q was also compared under standard reaction conditions (reaction

    time 18 h) to assess substituent effects (Table 4). Selective monoalkylation was

    observed in each case; however, the differences in reactivity were significant.

    Para-substituted anilines were more reactive than ortho- or metasubstituted

    analogues (Table 4, entry 1-2), wheras the reactivity of toluidines increased in the

    order ortho (1c) < meta (1d) < para (1e) probably owing to a combination of

    basicity and steric effects (Table 4, entry 4, 6 and 8). It was further more

    established that anilines comprising electron-withdrawing substituents were less

    reactive and that the reactivity of para-halogenated anilines decreased in the order:

    fluoro (1h), chloro (1i) > iodo (1k) (Table 4, entry 13, 15 and 19). Whereas para-

    substituted anilines bearing strong withdrawing groups including –COOCH3, -NO2,

    -CN do not give desired products (Table 4, entry 20-22).

    Table 4: Assessment of reactivity of functionalized anilines (1a-1q) in N-

    alkylation with 1-pentanol (2a).

    Entry 1 / R T

    [h]

    Temp.

    [°C]

    Conv.

    1 [%]a

    Sel. 3 [%]b

    1 1a p-OCH3 18 130 97 3aa 94 (91)

    2 1b o-OCH3 18 130 28 3ba 27

    3 1b o-OCH3 38 120 63 3ba 51 (42)

    4 1c o-CH3 18 130 23 3ca 12

    5 1c o-CH3 39 120 83 3ca 71(49)

    6 1d m-CH3 18 130 50 3da 40

    7 1d m-CH3 62 130 >99 3da >95 (84)

    8 1e p-CH3 18 130 88 3ea 75 (63)

    9 1e p-CH3 22 130 >99 3ea 90 (91)

    10 1f p-OH 18 130 >99 3fa 84 (69)

    11 1f p-OH 4 130 96 3fa 92 (94)

    12 1g o-F 18 130 25 3ga 13

    13 1h p-F 18 130 84 3ha 60

    14 1h p-F 42 120 >95 3ha 84(77)

    15 1i p-Cl 18 130 73 3ia 54

    16 1i p-Cl 63 120 87 3ia 77(76)

    17 1j p-Br, m-CH3 18 130 33 3ja 18

    18 1j p-Br, m-CH3 39 120 80 3ja 62(58)

    19 1k p-I 18 130 16 3ka 8

    20 1l p-COOCH3 18 130

  • Alkylation of amines with alcohols

    29

    Accordingly, the best result was obtained with p-hydroxyaniline, which was

    converted to 3fa in excellent 94% isolated yield already after 4 h (Table 4, entry

    11). This increased reactivity might be attributed to the enhanced nucleophilicity

    of 1f or the presence of the relatively acidic phenol moiety, which likely catalyzes

    the imine formation step. p-Methoxyaniline was converted to 3aa within 7 h (Table

    2), and para-methylaniline (1e), being less reactive, gave 3ea in 91% isolated

    yield after 22h (Table 4, entry 9). Unsubstituted aniline (1o) gave 3oa 90%

    isolated yield after 25 h (Table 4, entry 24).

    Other substrates required further optimization, which was initially conducted using

    one of the least reactive substrates 1b. It was shown that prolonged reaction times

    and the addition of molecular sieves (to accelerate the imine formation step) lead

    to satisfactory results and products 3ba (Table 4, entry 3), as well as 3ca, 3da,

    3ia and 3ja were isolated in moderate-to-high yields (42–84%) (Table 4, entry 5,

    7, 16 and 18). Taking advantage of the distinct difference in reactivity between

    substituted anilines, the selective monoalkylation of 1-methyl-2,4-diamino-

    benzene (1q) resulted in preferential formation of 3qa in 58% isolated yield (Table

    4, entry 26).

    Table 5: N-alkylation of p-methoxyaniline (1a) with various alcohols (2b-2l).

    Entry 2 Time [h] Conv.

    1a [%]a

    Sel. 3 [%]b Sel. 4 [%]b

    1 2b n-Octanol 18 85 3ab 79 (69) 4ab 6

    2 2c Ethanol 18 94 3ac 90 (85) 4ac 2

    3 2d Methanol 18 8 3ad 0 4ad 0

    4c 2e Propane-2-ol 24 42 3ae 12 4ae 24

    5c 2f Cyclohexanol 24 50 3af 14 4af 32

    6 2g Benzyl-alcohol 18 79 3ag 12 4ag 66

    7 2h Phenylethanol 18 93 3ah 87 (75) 4ah 5

    8 2i Ethane-1,2-diol 22 82 3ai 70 (74) 4ai 0

    9 2k Hexane-1,6-diol 22 83 3ak 65 (43) 4ak 0

    10d 2i Ethane-1,2-diol 24 80 3pi 40 4pi 0

    11d 2i Ethane-1,2-diol 42 90 3pi 60 (45) 4pi 0

    General reaction conditions: General Procedure A, 0.5 mmol 1a, 1 mmol 2a, 130 oC, 2 ml

    CPME, isolated yield in parenthesis. aConversion determined by GC-FID; bSelectivity

    determined by GC-FID; c2 mmol alcohol 2 was used; d3 mmol 2i was used and 2-

    aminoaniline (1p) was used instead of 1a. Main products also see Table 3.

    Monoalkylation of p-methoxyaniline with various alcohols Having

    established the reactivity pattern of anilines, we found that the selective

    monoalkylation of p-methoxyaniline (1a) proceeds with a variety of alcohols (2b-

    2l) with excellent results (Table 5). For instance, octanol (2b), ethanol (2c) as

    well as 2-phenyl-ethane-1-ol (2h) were readily converted to the corresponding

    amines 3ab, 3ac and 3ah in 69–85% isolated yields (Table 5, entry 1, 2 and 7).

  • Chapter 2

    30

    Alkylations using methanol were not successful (Table 5, entry 3), probably

    because the dehydrogenation of MeOH with employed catalytic system is

    unfavored. Interestingly, selective mono-alkylation with ethane-1,2-diol (2i) and

    hexane-1,6-diol (2k) delivered valuable amino-alcohols 3ai (74%) and 3ak (43%)

    (Table 5, entry 8-9). Notably, 2,3-dihydro-quinoxaline (3pi) containing a

    heterocyclic structure could also be constructed directly from inexpensive ethylene

    glycol (Table 3; Table 5, entry 10-11).

    N-alkylation of aliphatic amines with aliphatic alcohols It is important that

    aliphatic amines could also be successfully used as reaction partners with various

    alcohols (Figure 4). For instance, pentane-1-amine (5a) was alkylated with

    benzylalcohol (2g), providing 6ag in 67% yield. The same monoalkylated amine

    was obtained in 62% yield by a reverse route from benzylamine (7a) and 1-

    pentanol (2a), showing the flexibility and versatility of the new catalytic

    transformation. The reaction of piperidine (5b) with 2-phenylethylamine (2h)

    afforded tertiary amine 6bh in 53% isolated yield. Interestingly, furfuryl-amine

    (5c), which can be derived from lignocellulosic biomass through furfural, displayed

    an interesting reactivity towards bis-N-alkylation. In the reaction of 5c with 3 equiv

    of 2a, preferentially the corresponding tertiary amine (6caa) was formed.

    Figure 4: N-Alkylation of various aliphatic amines with alcohols. a Modular synthesis with

    aliphatic amines and alcohols. Pentyl-1-amine (5a) can be coupled with benzyl-alcohol (2g)

    or benzyl amine (7a) is coupled with 1-pentanol (2a) to afford the same amine product. b

    The methodology can be extended to secondary amines, piperidine (5b) is alkylated with

    2-phenyl-1-ethanol (2h). c The reaction of furfuryl-amine (5c) with 1-pentanol (2a)

    results in bis-N-alkylation product 6caa. aMolecular sieves were added.

  • Alkylation of amines with alcohols

    31

    N-alkylation of benzylamines N-alkylated benzyl amines are particularly

    important targets, as these moieties are present in a variety of drug molecules[29].

    Figure 5: Reactivity of various benzylamines (7) with 1-pentanol (2a) based on

    conversion in 6 h. General reaction conditions: General Procedure A, 0.5 mmol 7, 1 mmol

    2a, 130 oC, 2 ml CPME, conversion was determined by GC-FID using octadecane as internal

    standard.

    Again a distinct substituent effect was observed in the benzylamine reaction

    partner (Table 6). Para-methyl substituted 7g and unsubstituted 7a showed lower

    reactivity than meta-halogen substituted benzylamines (7b–7f), which reacted

    much faster with 1-pentanol (2a). This is probably due to the increased rate of

    reduction of the corresponding imines.[30] Substituted N-alkylated benzylamines

    8ba, 8ca, 8da and 8ea were obtained in excellent, 80–95% isolated yield.

    Following the alkylation over 6 h confirmed that the reactivity of benzyl amines

    7a–7e, increases in the order 7a

  • Chapter 2

    32

    Table 6: N-Alkylation of benzylamines with 1-pentanol and diols.

    General reaction conditions: General procedure A, 0.5 mmol 7, 1 mmol 2, 130 oC, 18-50

    h, isolated yields after purification. aSelectivity determined by GC-FID.

    Heterocycle formation with various benzylamines and diols Building on the

    excellent reactivity of benzyl amines 7c–7e, we attempted the formation of

    nitrogen-containing heterocycles of various sizes using diols of different chain

    lengths (butane-1,4-diol 2m, pentane-1,5-diol 2l or hexane- 1,6-diol 2k) (Table

    6). An additional advantage of the use of benzylamines is that the free N-

    heterocycles can be readily obtained by common debenzylation procedures.

    Despite the fact that a sequential catalytic N,N-dialkylation at the same nitrogen

    has to occur involving each of the hydroxyl groups of the diol, we reasoned that

    the first intermolecular alkylation is followed by an iminium ion-based alkylation,

    likely facilitated by the intramolecular nature of the second alkylation step. To our

    delight, pyrrolidine 9cm, containing a key five-membered heterocycle, was

    obtained in 60% yield from 7c and butane-1,4-ol (2m). Six-membered piperidine

    derivative 9cl was constructed using pentane-1,5 diol (2l) and 7c. As seven-

    membered ring formation is generally challenging in organic synthesis, it is a

    notable feature of our new Fe-based catalytic procedure that various azepane type

    N-heterocyclic compounds were readily obtained using benzyl amines 7b–7e and

    hexane-1,6-diol (2k). It is remarkable that in all these reactions, full conversion

    and perfect product selectivity was observed. Among all derivatives, the chloro-

    substituted 9bk was obtained with the highest isolated yield (85%). In contrast,

  • Alkylation of amines with alcohols

    33

    the reaction of unsubstituted benzylamine 7a with 2k afforded the monoalkylation

    product, showing that the second cyclization step was much slower in this case.

    These results highlight the value of this methodology in accessing biological

    important nitrogen heterocycles, which contain -F and -CF3 substituents. These are

    frequently encountered motives in pharmaceutically active compounds (Figure 6).

    Synthesis of pharmaceutically relevant molecules To demonstrate the power

    of the new C–N bond formation, we applied the Fe-catalyzed direct amine

    alkylation as a key connection step in the synthesis of the N-arylpiperazine drug

    Piribedil (12) (Figure 6), a dopamine antagonist used in the treatment of

    Parkinson’s disease[11,31]. The application of iron-based homogeneous catalyst in

    the preparation of pharmaceutically active molecules is highly desired, given the

    limited allowed ppm levels of noble metal catalysts in the final products[29].

    Commercially available 1-(2-pyrimidyl)-piperazine (10) and 4 equiv of piperonyl

    alcohol (11) were used as the substrates, providing Piribedil in 54% isolated yield.

    This direct, base free catalytic coupling illustrates several advantages for drug

    syntheses: no auxiliary reagents are required, an iron catalyst and the green

    solvent CPME are used, and water is the only waste product.

    Figure 6: Direct synthesis of Piribedil 12.

    Conclusion

    We have developed a versatile Fe-based catalytic approach for the direct coupling

    of amines and alcohols. Key elements are a single bifunctional Fe-catalyst for

    alcohol dehydrogenation and imine hydrogenation, and the use of the green

    solvent CPME. The variety of primary amines and alcohols, including those derived

    from biomass, which can be coupled, provides an excellent basis for highly flexible

    methodology to synthesize a broad array of secondary amines and nitrogen

    heterocycles and holds promise for future application in diversity-oriented

    synthesis strategies. The first upscaling experiments showed that this direct N-

    alkylation can provide up to gram amounts of secondary amines. Mechanistic

    investigations are needed to elucidate the role of steric and electronic parameters

    in the rate-determining step, which can be both the imine formation as well as the

    imine reduction. For example, p-hydroxy aniline was rapidly converted within 4 h,

    whereas o-methoxy aniline was much less reactive. Similarly, benzyl-amines

    display marked reactivity difference, depending on the nature and position of the

    substituents at the aromatic ring. We foresee that control of the reactivity of well-

    defined iron catalysts, as shown here successfully for the borrowing hydrogen

    strategy in the direct alkylation of amines with alcohols, will enable the discovery

    of a range of other new catalyst structures and sustainable transformations, using

    abundant iron.

  • Chapter 2

    34

    Experimental section

    General methods

    Chromatography: Merck silica gel type 9385 230-400 mesh or Merck Al2O3 90

    active neutral, TLC: Merck silica gel 60, 0.25 mm or Al2O3 60 F254 neutral.

    Components were visualized by UV, Ninhydrin or I2 staining. Progress of the

    reactions was determined by GC-MS (GC: HP 6890, MS: HP 5973) with an HP012

    column (Agilent Technologies, Palo Alto, CA). Mass spectra were recorded on an

    AEI-MS-902 mass spectrometer (EI+) or a LTQ Orbitrap XL (ESI+). Conversions

    were determined by GC-FID (GC: HP 6890) with an HP-5 column (Agilent

    Technologies, Palo Alto, CA). GC-MS and GC-FID analysis method: 60 °C 5 min,

    180 °C 5 min (10 °C/min), 260 oC 5 min (10 °C/min). 1H- and 13C NMR spectra

    were recorded on a Varian AMX400 (400 and 100.59 MHz, respectively) using

    CDCl3 or CD2Cl2 as solvent. In situ 1H NMR spectra were recorded on a Varian Unity

    Plus Varian-500 (500 MHz) using toluene-d8 as solvent. Chemical shift values are

    reported in ppm with the solvent resonance as the internal standard (CDCl3: 7.26

    for 1H, 77.00 for 13C; CD2Cl2: 5.32 for 1H, 53.84 for 13C; toluene-d8: 7.09 for 1H).

    Data are reported as follows: chemical shifts, multiplicity (s = singlet, d = doublet,

    t = triplet, q = quartet, br = broad, m = multiplet), coupling constants (Hz), and

    integration. All reactions were carried out under an Argon atmosphere using oven

    dried glassware and using standard Schlenk techniques. THF and toluene were

    collected from a MBRAUN solvent purification system (MB SPS-800). Dioxane

    (99.5%, extra dry), dichloroethane (DCE, 99.8%, extra dry), N,N-

    dimethylformamide (DMF, 99.8%, extra dry) and acetonitrile (CH3CN, 99.9%,

    extra dry) were purchased from Acros without further purification. Aniline, 4-

    methylbenzylamine, benzylamine, 3-trifluoromethylbenzylamine were purified by

    distillation. 4-Methylaniline was purified by recrystallization. 2-Methoxyaniline was

    purified by column chromatography. All other reagents were purchased from

    Sigma or Acros in reagent or higher grade and were used without further

    purification. Complex Cat 3, Cat 3b was synthetized according to literature

    procedures[32].

    Preparation of Cat 3 and Cat 3b

    1,8-Bis(trimethylsilyl)-1,7-octadiyne To a solution of 1,7-octadiyne (2 mL,

    1.56 g, 15 mmol) in THF (20 mL), a solution of n-BuLi (20.6 mL of a 1.6M in

    hexane solution, 33 mmol, 2.2 equiv) was added dropwise over 20 min at -78 °C

    under N2, in a 100ml Schlenk tube. After stiring for 10 min at -78 °C, and 1 h

    under room temperature, trimethylsilyl chloride (8.4 ml, 66 mmol, 4.4 equiv) was

    added slowly and the mixture was stirred at room temperature for 16 hours. The

    cloudy, white mixture was quenched with saturated aqueous NH4Cl (5 mL) followed

    by water until all of the white precipitate dissolved. The organic layer was removed

  • Alkylation of amines with alcohols

    35

    and the aqueous layer was extracted with pentane. The combined organic layers

    were dried over anhydrous Na2SO4 and the solvent was evaporated under reduced

    pressure to give a yellow oil. Purification of the crude product by Kugelrohr

    distillation afforded 3.56 g (95% yield) of a light yellow oil/solid mixture (product

    is a low-melting solid). 1H NMR (400 MHz, CDCl3, ppm): δ 2.23–2.28 (m, 4H),

    1.60–1.64 (m, 4H), 0.15 (s, 18H). The physical data were identical in all respects

    to those previously reported.[32]

    Tricarbonyl(1,3-bis(trimethylsilyl)-4,5,6,7-tetrahydro-2H-inden-2-

    one)iron (Cat 3) A solution of 1,8-bis(trimethylsilyl)-1,7-octadiyne (1.2 g, 4.8

    mmol) and diiron nonacarbonyl (1.74 g, 4.8 mmol, 1 equiv) in toluene (60 mL) in

    a 250 ml oven-dried Schlenk tube was heated to 120 °C and stirred for 24 hours

    under argon. After cooling, the mixture was filtered over celite. Then the solution

    was concentrated and filtered over silica to remove all the iron black. The solution

    was evaporated under reduced pressure. The remaining solid residue was washed

    with cold pentane and crystalized in heptane affording yellow a crystalline solid,

    which is air and moisture stable (1.36g, 68% yield). 1H NMR (400 MHz, CDCl3): δ

    2.50–2.62 (m, 4H), 1.77–1.87 (m, 4H), 0.26 (s, 18 H). 13C NMR (100 MHz, CDCl3):

    δ 209.04, 181.22, 110.99, 71.74, 24.77, 22.41, –0.28. The physical data were

    identical in all respects to those previously reported.[32]

    Acetonitrile dicarbonyl(1,3-bis(trimethylsilyl)-4,5,6,7-tetrahydro-2H-

    inden-2-one)iron (Cat 3b) An oven-dried 250 ml Schlenk tube was charged with

    100 ml dry acetone and 2 ml dry CH3CN and degassed with N2 for 20 min. Then 1

    g Cat 3 (2.38 mmol) was added under N2, stirring for 1 min until it fully solubilized.

    216 mg Me3NO (1.2eq) was added under N2. A direct color change from yellow to

    orange was observed in 5 s. The conversion of Cat 3 can be monitored by TLC

    (pentane/ethyl acetate = 1/1, on silica gel, RfCat1a = 0.95, RfCat1b = 0.35). After 1

    h, the solvent was removed under a vacuum; Cat 3b was purified through flash

    chromatography, and obtained as brown solid (0.91g, 88% yield). 1H NMR (400

    MHz, CDCl3): δ 2.05 – 2.48 (m, 4H), 2.21 (s, 3H), 1.38 – 1.73 (m, 4H), 0.22 (s,

    18 H). 13C NMR (100 MHz, CDCl3): δ 212.80, 180.12, 126.00, 106.58, 69.91, 24.83,

    22.31, 4.43, -0.12. The physical data were identical in all respects to those

    previously reported.[32]

    Cat 3 and Cat 3b are slightly light sensitive but air stable.

    Representative procedures

    General procedure A: An oven-dried 20 ml Schlenk tube, equipped with a stirring

    bar, was charged with amine (0.500 mmol, 1 equiv), alcohol (given amount), iron

    complex Cat 3 (0.025 mmol, 10.5 mg), Me3NO (0.050 mmol, 3.75 mg) and

    cyclopentyl methyl ether (solvent, 2 ml). Solid materials were weighed into the

    Schlenk tube under air, the Schlenk tube was subsequently connected to an argon

    line and vacuum-argon exchange was performed three times. Liquid starting

    materials and solvent were charged under an argon stream. The Schlenk tube was

    capped and the mixture was rapidly stirred at room temperature for 1 min, then it

    was placed in a pre-heated oil bath at the appropriate temperature and stirred for

  • Chapter 2

    36

    a given time. The reaction mixture was cooled down to room temperature and the

    crude mixture was filtered through celite, eluted with ethyl-acetate, and

    concentrated in vacuo. The residue was purified by flash column chromatography

    to provide the pure amine product.

    General procedure B (With molecular sieves): An oven-dried 20 ml Schlenk tube,

    equipped with stirring bar, was charged with amine (0.5 mmol, 1 equiv), alcohol

    (given amount), iron complex Cat 3 (0.025 mmol, 10.5 mg), Me3NO (0.05 mmol,

    3.75 mg) and cyclopentyl methyl ether (solvent, 2 ml). The solid starting materials

    were added into the Schlenk tube under air, the Schlenk tube was subsequently

    connected to an argon line and a vacuum-argon exchange was performed three

    times. Liquid starting materials and solvent were charged under an argon stream

    followed by addition of 35.0–45.0 mg activated molecular sieves 4A. The Schlenk

    tube was capped and the mixture was rapidly stirred at room temperature for 1

    minute, then was placed into a pre-heated oil bath at the appropriate temperature

    and stirred for a given time. The reaction mixture was cooled down to room

    temperature and the crude mixture was filtered through celite, eluted with ethyl

    acetate, and concentrated in vacuo. The residue was purified by flash column

    chromatography to provide the pure amine product.

    Upscaling procedure for N-alkylation of p-methoxyaniline An oven-dried

    100 ml Schlenk tube, equipped with a stirring bar, was charged with p-

    methoxyaniline (7.40 mmol, 0.910 g, 1 equiv), iron complex Cat 3 (0.37 mmol,

    155 mg, 0.05 equiv) and Me3NO (0.74 mmol, 55.5 mg, 0.10 equiv) under air and

    the Schlenk tube was subsequently connected to an argon line and vacuum-argon

    exchange was performed three times. Then 1-pentanol (14.8 mmol, 1.62 ml, 2

    equiv), cyclopentylmethylether (solvent, 30 ml), 3Å molecular sieves (800 mg)

    were charged under an argon stream. The Schlenk tube was capped and the

    mixture was rapidly stirred at room temperature for 3 min, then was placed into a

    pre-heated oil bath at 130 °C and stirred for 18 h. After cooling down to room

    temperature, the crude mixture was concentrated in vacuo. The residue was

    purified by flash column chromatography (SiO2, n-pentane/EtOAc 100:0 to 95:5)

    to provide 1.085 gram of p-methoxy-N-pentylaniline in 76% isolated yield.

    General procedure for in situ 1H NMR study of the N-alkylation of p-

    methoxyaniline (1a) with 1-pentanol (2a) in toluene-d8 at 100 °C. 0.12 mmol p-

    anisidine (1a), 0.24 mmol 1-pentanol (2a), 0.012 mmol Cat 3, 0.024 mmol Me3NO

    and 0.6 ml toluene-d8 were added in a J-Young NMR tube under argon. The tube

    was sealed in an argon flow and placed in the pre-heated (100 °C) NMR machine,

    shimming was performed at reaction temperature. In the following 8 h, 16 spectra

    were recorded.

  • Alkylation of amines with alcohols

    37

    Spectral data of isolated compounds

    4-methoxy-N-pentylaniline (3aa): Synthesized according to

    General procedure A. p-Anisidine (0.062 g, 0.500 mmol) affords

    3aa (0.088 g, 91% yield). Yellow oil obtained after column

    chromatography (SiO2, N-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 6.78 (d, J = 8.9 Hz, 2H), 6.59 (d, J = 8.8 Hz,

    2H), 3.75 (s, 3H), 3.06 (t, J = 7.2 Hz, 2H), 1.55 - 1.69 (m, 2H),

    1.30 - 1.50 (m, 4H), 0.92 (t, J = 7.0 Hz, 3H). 13C NMR (100 MHz,

    CDCl3) δ 151.93, 142.79, 114.85, 114.00, 55.78, 44.99, 29.34, 22.50, 14.01. The

    physical data were identical in all respects to those previously reported.[33]

    2-methoxyl-N-pentylaniline (3ba): Synthesized according to

    General procedure B. o-Anisidine (0.056 ml, 0.5 mmol) affords

    3ba (0.041 g, 42% yield). Yellow oil obtained after column

    chromatography (SiO2, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 6.90 (td, J = 7.6, 1.1 Hz, 1H), 6.79 (d, J =

    7.9 Hz, 1H), 6.58 – 6.74 (m ,2H), 4.05 – 4.35 (br.s, 1H), 3.87

    (s, 3H), 3.14 (t, J = 7.1 Hz, 2H), 1.61 – 1.75 (m, 2H), 1.32 –

    1.50 (m, 4H), 0.95 (t, J = 6.8 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ 146.68, 138.47,

    121.26, 116.02, 109.69, 109.29, 55.33, 43.67, 29.40, 29.23, 22.52, 14.02. HRMS

    (APCI+, m/z): calculated for C12H20NO [M+H]+:194.15394; found: 194.15387.[34]

    2-methyl-N-pentylaniline (3ca): Synthesized according to

    General procedure B. o-Toluidine (0.053 ml, 0.5 mmol) affords 3ca

    (0.043 g, 49% yield). Yellow oil obtained after column

    chromatography (SiO2, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 7.15 (m, J = 7.7 Hz, 1H), 7.07 (d, J = 7.2, 1H),

    6.60 – 6.70 (m, 2H), 3.35 – 3.60 (br.s, 1H), 3.17 (t, J = 7.1 Hz,

    2H), 2.16 (s, 3H), 1.64 – 1.74 (m, 2H), 1.35 – 1.50 (m, 4H), 0.96

    (t, J = 6.9 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ 146.36, 129.96,

    127.10, 121.63, 116.58, 109.58, 43.93, 29.41, 29.30, 22.52, 17.42, 14.04. HRMS

    (APCI+, m/z): calculated for C12H20N [M+H]+: 178.15903; found: 178.15896.[34]

    3-methyl-N-pentylaniline (3da): Synthesized according to

    General procedure A. m-Toluidine (0.054 ml, 0.5 mmol) affords

    3da (0.074 g, 84% yield). Yellow oil obtained after column

    chromatography (SiO2, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 7.05 – 7.15 (m, 1H), 6.51 – 6.62 (m, 1H),

    6.38 – 6.51 (m, 2H), 3.32 – 3.85 (br.s, 1H), 3.13 (t, J = 7.2 Hz,

    2H), 2.32 (s, 3H), 1.57 – 1.72 (m, 2H), 1.32 – 1.52 (m, 4H), 0.96

    (t, J = 7.1 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ 148.52, 138.89, 129.03, 117.97,

    113.42, 109.82, 43.95, 29.32, 29.26, 22.48, 21.59, 14.02. HRMS (APCI+, m/z):

    calculated for C12H19N [M+H]+: 178.15903; found: 178.15874.[35]

  • Chapter 2

    38

    4-methyl-N-pentylaniline (3ea): Synthesized according to

    General procedure A. p-Toluidine (0.054 g, 0.5 mmol) affords 3ea

    (0.081 g, 91% yield). Yellow oil obtained after column

    chromatography (SiO2, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 6.99 (d, J = 8.1 Hz, 2H), 6.56 (d, J = 8.2 Hz,

    2H), 3.08 (t, J = 7.1 Hz, 2H), 2.24, (s, 3H), 1.56 – 1.67 (m, 2H),

    1.27 – 1.50 (m, 4H), 0.91 (t, J = 6.8 Hz, 3H). 13C NMR (100 MHz,

    CDCl3) δ 146.26, 129.65, 126.23, 112.87, 44.34, 29.34, 29.30, 22.50, 20.33,

    14.02.[33]

    4-(pentylamino)phenol (3fa): Synthesized according to General

    procedure A. 4-Aminophenol (0.055 g, 0.5 mmol) affords 3fa (0.084

    g, 94% yield).Yellow oil obtained after column chromatography

    (SiO2, n-pentane/EtOAc 90:10 to 60:40). 1H NMR (400 MHz, CDCl3)

    δ 6.68 (d, J = 8.0 Hz, 2H), 6.55 (d, J = 8.2 Hz, 2H), 3.90 – 1.69

    (br.s, 2H), 3.05 (t, J = 6.8 Hz, 2H), 1.52 – 1.69 (m, 2H), 1.28 –

    1.45 (m, 4H), 0,91 (t, J = 7.0 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ

    148.12, 142.23, 116.23, 114.83, 45.42, 29.30, 22.47, 13.99. HRMS

    (APCI+, m/z): calculated for C11H25N [M+H]+: 180.13829; found: 180.13823.

    4-fluoro-N-pentylaniline (3ha): Synthesized according to

    General procedure A. 4-fluoroaniline (0.047 ml, 0.5 mmol) affords

    3ha (0.070 g, 77% yield). Yellow oil obtained after column

    chromatography (SiO2, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 6.80 – 6.95 (m, 2H), 6.47 – 6.58 (m, 2H), 3.27

    – 3.59 (br.s, 1H), 3.06 (t, J = 7.1 Hz, 2H), 1.55 - 1.67 (m, 2H),

    1.31 - 1.44 (m, 4H), 0.88 – 0.99 (m, 3H). 13C NMR (100 MHz,

    CDCl3) δ 155.61 (d, J = 234.4 Hz), 144.86, 115.53 (d, J = 22.3 Hz), 113.49 (d, J

    = 7.4 Hz), 44.63, 29.30, 29.20, 22.48, 14.00.[36]

    4-Chloro-N-pentylaniline (3ia): Synthesized according to

    General procedure A. 4-Chloroaniline (0.063 g, 0.5 mmol)

    affords 3ia (0.075 g, 76% yield). Yellow oil obtained after column

    chromatography (SiO2, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 7.11 (d, J = 8.6 Hz, 2H), 6.54 (d, J = 8.7 Hz,

    2H), 3.75 (s, 3H), 3.07 (t, J = 7.2 Hz, 2H), 1.55 - 1.67 (m, 2H),

    1.27 - 1.44 (m, 4H), 0.92 (t, J = 6.7 Hz, 3H). 13C NMR (100

    MHz, CDCl3) δ 147.02, 128.93, 121.45, 113.63, 44.03, 29.25,

    29.08, 22.45, 13.98.[33]

    4-bromo-3-methyl-N-pentylaniline (3ja): Synthesized

    according to General procedure B. 4-Bromo-3-methylaniline

    (0.092 g, 0.5 mmol) affords 3ja (0.074 g, 58% yield). Yellow

    oil obtained after column chromatography (SiO2, n-

    pentane/EtOAc 100:0 to 95:5). 1H NMR (400 MHz, CDCl3) δ

    7.26 (d, J = 8.6 Hz, 1H), 6.48 (d, J = 2.8 Hz, 1H), 6.31 (dd, J

    = 8.6, 2.8 Hz, 1H), 3.35 – 3.80 (br.s, 1H), 3.06 (t, J = 7.1 Hz,

    2H), 2.31 (s, 3H), 1.50 – 1.65 (m, 2H), 1.30 – 1.42 (m, 4H), 0.92 (t, J = 6.8 Hz,

  • Alkylation of amines with alcohols

    39

    3H). 13C NMR (100 MHz, CDCl3) δ 147.74, 138.14, 132.56, 114.93, 111.83, 111.32,

    43.99, 29.27, 29.12, 23.05, 22.47, 14.01. HRMS (APCI+, m/z): calculated for

    C12H19NBr [M+H]+: 258.06954; found: 258.06738.

    N-Pentylaniline (3oa): Synthesized according to General

    procedure A. Aniline (0.046 ml, 0.5 mmol) affords 3oa (0.073 g,

    90% yield). Yellow oil obtained after column chromatography (SiO2,

    n-pentane/EtOAc 100:0 to 95:5). 1H NMR (400 MHz, CDCl3) δ 7.15

    – 7.25 (m, 2H), 6.68 – 6.67 (m, 1H), 6.59 – 6.67 (m, 2H), 3.45 –

    3.80 (br.s, 1H), 3.13 (t, J = 7.2 Hz, 2H), 1.57 – 1.71 (m, 2H), 1.32

    – 1.48 (m, 4H), 0.95 (t, J = 7.0 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ 148.45,

    129.16, 117.03, 112.65, 43.94, 29.31, 29.22, 22.49, 14.02. The physical data

    were identical in all respects to those previously reported.[33]

    4-Methyl-N1-pentylbenzene-1,3-diamine (3qa): Synthesized

    according to General procedure A. 2,4-Diaminotoluene (0.061 g, 0.5

    mmol) affords 3qa (0.056 g, 58% yield). Yellow oil obtained after

    column chromatography (SiO2, n-pentane/EtOAc 90:10 to 50:50). 1H

    NMR (400 MHz, CDCl3) δ 6.85 (d, J = 8.0, 1H), 6.05 (dd, J = 8.0, 2.2

    Hz, 1H), 6.00 (d, J = 2.2 Hz, 1H), 3.25 – 3.65 (br.s, 3H), 3.07 (t, J =

    7.2 Hz, 2H), 2.08 (s, 3H), 1.55 – 1.66 (m, 2H), 1.33 – 1.43 (m, 4H),

    0.94 (t, J = 7.0, 3H). 13C NMR (100 MHz, CDCl3) δ 147.93, 145.17,

    130.91, 111.44, 103.87, 99.64, 44.21, 29.29, 29.27, 22.46, 16.33, 14.00. NOESY-

    NMR spectrum see figure 7. HRMS (APCI+, m/z): calculated for C12H21N2 [M+H]+:

    193.16993; found: 193.16962.

    1,2,3,4-Tetrahydroquinoxaline (3pi): Synthesized according to

    General procedure A. o-Phenylenediamine (0.054 g, 0.5 mmol) affords

    3pi (0.030 g, 45% yield). Orange solid obtained after column

    chromatography (SiO2, n-pentane/EtOAc 90:10 to 50:50). 1H NMR (400

    MHz, CDCl3) δ 6.55 – 6.63 (m, 2H), 6.46 – 6.54 (m, 2H), 3.42 (s, 4H), 3.25 –

    3.54 (br.s, 2H). 13C NMR (100 MHz, CDCl3) δ 133.57, 118.77, 114.72, 41.34. The

    physical data were identical in all respects to those previously reported.[37]

    4-Methoxy-N-octylaniline (3ab): Synthesized according

    to General procedure A. p-Anisidine (0.062 g, 0.5 mmol)

    affords 3ab (0.081 g, 69% yield). Yellow oil obtained after

    column chromatography (SiO2, n-pentane/EtOAc 100:0 to

    95:5). 1H NMR (400 MHz, CDCl3) δ 6.68 (d, J = 8.7 Hz, 2H),

    6.60 (d, J = 8.7 Hz, 2H), 3.75 (s, 3H), 3.06 (t, J = 7.2 Hz,

    2H), 1.55 – 1.65 (m, 2H), 1.34 – 1.44 (m, 2H), 1.16 – 1.44

    (m, 8H). 13C NMR (100 MHz, CDCl3) δ 151.96, 142.81, 114.88, 114.03, 55.81,

    45.05, 31.81, 29.67, 29.42, 29.25, 27.19, 22.64, 14.08. The physical data were

    identical in all respects to those previously reported.[38]

  • Chapter 2

    40

    4-Methoxy-N-ethylaniline (3ac): Synthesized according to General

    procedure A. p-Anisidine (0.062 g, 0.5 mmol) affords 3ac (0.064 g, 85%

    yield). Yellow oil obtained after column chromatography (SiO2, n-

    pentane/EtOAc 100:0 to 95:5). 1H NMR (400 MHz, CDCl3) δ 6.79 (d, J =

    7.1 Hz, 2H), 6.60 (d, J = 7.3 Hz, 2H), 3.75 (s, 3H), 3.02 – 3.23 (m, 2H),

    1.24 (t, J = 6.8, 3H). 13C NMR (100 MHz, CDCl3) δ 152.04, 142.70, 114.86,

    114.10, 55.79, 39.44, 14.97. The physical data were identical in all respects to

    those previously reported.[39]

    4-Methoxy-N-phenethylaniline (3ah): Synthesized according to

    General procedure A. p-Anisidine (0.062 g, 0.5 mmol) affords 3ah

    (0.085 g, 75% yield). Yellow oil obtained after column

    chromatography (SiO2, n-pentane/EtOAc 100:0 to 95:5).1H NMR

    (400 MHz, CDCl3) δ 7.28 - 7.46 (m, 2H), δ 7.12 – 7.28 (m, 3H),

    6.79 (d, J = 8.8 Hz, 2H), 6.61 (d, J = 8.6 Hz, 2H), 3.75 (s, 3H),

    3.37 (t, J = 7.0 Hz, 2H), 2.91 (t, J = 6.9 Hz, 2H). 13C NMR (100

    MHz, CDCl3) δ 152.15, 142.16, 139.35, 128.73, 128.52, 126.32,

    114.88, 114.36, 55.74, 46.01, 35.54. The physical data were identical in all

    respects to those previously reported.[40]

    2-((4-Methoxyphenyl)amino)ethanol (3ai): Synthesized

    according to General procedure A. p-Anisidine (0.062 g, 0.5 mmol)

    affords 3ai (0.062 g, 74% yield). Yellow oil obtained after column

    chromatography (SiO2, n-pentane/EtOAc 90:10 to 50:50). 1H NMR

    (400 MHz, CDCl3) δ 6.74 – 6.84 (m, 2H), 6.58 – 6.66 (m, 2H), 3.79

    (t, J = 5.2 Hz, 2H), 3.75 (s, 3H), 3.23 (t, J = 5.2 Hz, 2H), 2.86 – 3.08

    (br.s, 1H). 13C NMR (100 MHz, CDCl3) δ 152.49, 142.14, 114.86,

    114.76, 61.18, 55.75, 47.16. The physical data were identical in all respects to

    those previously reported.[41]

    6-((4-Methoxyphenyl)amino)hexan-1-ol (3ak):

    Synthesized according to General procedure A. p-Anisidine

    (0.062 g, 0.5 mmol) affords 3ak (0.048 g, 43% yield). Yellow

    oil obtained after column chromatography (SiO2, n-

    pentane/EtOAc 90:10 to 50:50). 1H NMR (400 MHz, CDCl3) δ

    6.72 – 6.82 (m, 2H), 6.52 – 6.62 (m, 2H), 3.74 (s, 3H), 3.63

    (t, J = 6.6 Hz, 2H), 3.06 (t, J = 7.1 Hz, 2H), 2.31 – 2.51 (br.s,

    2H), 1.51 – 1.68 (m, 4H), 1.31 – 1.43 (m, 4H). 13C NMR (100

    MHz, CDCl3) δ 151.99, 142.64, 114.85, 114.10, 62.75, 55.79, 44.94, 32.61, 29.56,

    26.92, 25.55. HRMS (APCI+, m/z): calculated for C13H21NO2 [M+H]+: 224.16451;

    found: 224.16428.

    N,N-Dipentylfurfurylamine (6caa): Synthesized

    according to General procedure B. Furfurylamine (0.044 ml,

    0.5 mmol) affords 6caa (0.075 g, 63% yield). Yellow oil

    obtained after column chromatography (SiO2, n-

    pentane/EtOAc 95:5 to 60:40). 1H NMR (400 MHz, CDCl3) δ

    7.35 (d, J = 2.0, 1H), 6.30 (dd, J = 2.8, 2.0 Hz, 1H), 6.15 (d, J = 3.0, 1H), 3.64

  • Alkylation of amines with alcohols

    41

    (s, 2H), 2.40 (m, J = 7.6, 2H), 1.40 – 1.62 (m, 4H), 1.15 – 1.40 (m, 8H), 0.88

    (t, J = 7.0, 6H). 13C NMR (100 MHz, CDCl3) δ 152.88, 141.66, 109.88, 108.11,

    53.84, 49.97, 29.71, 26.74, 22.60, 14.06. HRMS (APCI+, m/z): calculated for

    C15H25NO [M+H]+: 238.21666; found: 238.21654.

    1-Phenethylpiperidine (6bh): Synthesized according to General

    procedure B. Piperidine (0.049 ml, 0.5 mmol) affords 6bh (0.050 g, 52%

    yield). Yellow oil obtained after column chromatography (Al3O2, n-

    pentane/EtOAc 100:0 to 90:10). 1H NMR (400 MHz, CD2Cl2) δ 7.23 –

    7.34 (m, 2H), 7.10 – 7.34 (m, 3H), 2.72 – 2.80 (m, 2H), 2.48 – 2.56

    (m, 2H), 2.35 – 2.48 (m, 4H), 1.52 – 1.62 (m, 4H), 1.33 – 1.48 (m, 2H). 13C NMR (100 MHz, CDCl3) δ 140.60, 128.66, 128.29, 125.88, 61.43,

    54.52, 33.62, 25.97, 24.39. The physical data were identical in all respects to

    those previously reported.[42]

    N-Pentyl-phenethylamine (6ah or 6da): Synthesized

    according to General procedure A. Phenethylamine (0.063 ml,

    0.5 mmol) affords 6da (0.031 g, 32% yield). Amylamine (0.058

    ml, 0.5 mmol) affords 6ah (0.014 g, 15% yield). Yellow oil

    obtained after column chromatography (SiO2, MeOH/EtOAc

    0:100 to 5:95). 1H NMR (400 MHz, CD2Cl2) δ 7.24 – 7.33 (m,

    2H), 7.12 – 7.24 (m, 3H), 2.79 – 2.91 (m, 2H), 2.68 – 2.79 (m, 2H), 2.58 (t, J =

    7.2 Hz), 1.38 – 1.48 (m, 2H), 1.22 – 1.32 (m, 4H), 0.89 (t, J = 7.0 Hz, 3H). 13C

    NMR (100 MHz, CD2Cl2) δ 141.03, 129.09, 128.68, 126.29, 51.65, 50.19, 36.83,

    30.21, 29.98, 23.03, 14.24. The physical data were identical in all respects to

    those previously reported.[43]

    N-Pentylbenzylamine (6ag or 8aa): Synthesized according

    to General procedure A. Benzylamine (0.055 ml, 0.5 mmol)

    affords 8aa (0.055 g, 62% yield). Amylamine (0.058 ml, 0.5

    mmol) affords 6ag (0.059 g, 67% yield). Yellow oil obtained

    after column chromatography (SiO2, n-pentane/EtOAc 50:50 to

    0:100). 1H NMR (400 MHz, CD2Cl2) δ 7.26 – 7.37 (m, 4H), 7.16

    – 7.26 (m, 1H), 3.77 (s, 2H), 2.61 (t, J = 7.2 Hz, 2H), 1.57 – 1.80 (br.s, 1H), 1.45

    – 1.55 (m, 2H), 1.27 – 1.36 (m, 4H), 0.91 (t, J = 7.0 Hz, 3H). 13C NMR (100 MHz,

    CD2Cl2) δ 141.46, 128.62, 128.45, 127.07, 54.33, 49.89, 30.24, 30.01, 23.07,

    14.27. The physical data were identical in all respects to those previously

    reported.[43]

    3-Chloro-N-pentylbenzylamine (8ba): Synthesized

    according to General procedure A. 3-Chlorobenzylamine (0.061

    ml, 0.5 mmol) affords 8ba (0.084 g, 80% yield). Yellow oil

    obtained after column chromatography (SiO2, n-pentane/EtOAc

    50:50 to 0:100). 1H NMR (400 MHz, CDCl3) δ 7.32 (s, 1H), 7.06

    – 7.25 (m, 3H), 3.76 (s, 2H), 2.60 (t, J = 7.2 Hz, 2H), 1.59 – 1.70 (br.s, 1H), 1.45

    – 1.55 (m, 2H), 1.27 – 1.35 (m, 4H), 0.89 (t, J = 6.9 Hz, 3H). 13C NMR (100 MHz,

    CDCl3) δ 142.50, 134.17, 129.56, 128.14, 126.98, 126.16, 53.40, 49.37, 29.67,

  • Chapter 2

    42

    29.46, 22.56, 14.01. HRMS (APCI+, m/z): calculated for C12H19ClN [M+H]+:

    212.12005; found: 212.11983.[44]

    3-Fluoro-N-pentylbenzylamine (8ca): Synthesized

    according to General procedure A. 3-Fluorobenzylamine (0.057

    ml, 0.5 mmol) affords 8ca (0.085 g, 87% yield). Yellow oil

    obtained after column chromatography (SiO2, n-pentane/EtOAc

    50:50 to 0:100). 1H NMR (400 MHz, CDCl3) δ 7.20 – 7.34 (m,

    1H), 6.98 – 7.18 (M, 2H), 6.93 (td, J = 8.4, 2.1 Hz, 1H), 3.79

    (s, 2H), 2.61 (t, J = 7.3 Hz, 2H), 1.76 – 2.04 (br.s, 1H), 1.45 – 1.61 (m, 2H), 1.22

    – 1.37 (m, 4H), 0.89 (t, J = 6.9 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ 162.93 (d,

    J = 245.5 Hz), 142.96 (d, J = 7.3 Hz), 129.73 (d, J = 8.2 Hz), 129.63 (d, J = 2.8

    Hz), 114.86 (d, J = 21.2 Hz), 113.70 (d, J = 21.2 Hz), 53.35, 49.31, 29.62, 29.47,

    22.56, 14.00. HRMS (APCI+, m/z): calculated for C12H19FN [M+H]+: 196.14960;

    found: 196.14938.[44]

    3-Trifluoromethyl-N-pentylbenzylamine (8da):

    Synthesized according to General procedure A. 3-

    Trifluorobenzylamine (0.072 ml, 0.5 mmol) affords 8da (0.115

    g, 94% yield). Yellow oil obtained after column chromatography

    (SiO2, n-pentane/EtOAc 50:50 to 0:100). 1H NMR (400 MHz,

    CDCl3) δ 7.60 (s, 1H), 7.46 – 7.56 (m, 2H), 7.37 – 7.46 (m,

    1H), 3.84 (s, 2H), 2.62 (t, J = 7.2 Hz, 2H), 2.06 – 2.21 (br.s, 1H), 1.45 – 1.57 (m,

    2H), 1.26 – 1.35 (m, 4H), 0.89 (t, J = 6.8 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ

    141.31, 131.44, 130.66 (q, J = 32.3 Hz), 128.72, 124.79 (q, J = 3.7 Hz), 124.20

    (d, J = 272.2 Hz), 123.73 (q, J = 3.8 Hz), 53.41, 49.39, 29.61, 29.46, 22.55,

    13.98. HRMS (APCI+, m/z): calculated for C12H29F3N [M+H]+: 246.14641; found:

    246.14594.[44]

    3-Fluoro-5-trifluoromethyl-N-pentylbenzylamine

    (8ea): Synthesized according to General procedure A. 3-

    Fluoro-5-trifluoromethylbenzylamine (0.073 ml, 0.5

    mmol) affords 8ea (0.125 g, 95% yield). Yellow oil

    obtained after column chromatography (SiO2, n-

    pentane/EtOAc 50:50 to 0:100). 1H NMR (400 MHz, CDCl3) δ 7.40 (s, 1H), 7.28

    (d, J = 9.1 Hz, 1H), 7.19 (d, J = 8.0 Hz, 1H), 3.88 – 4.04 (br.s, 1H), 3.48 (s, 2H),

    2.61 (t, J = 7.1 Hz, 2H), 1.43 – 1.59 (m, 2H), 1.25 – 1.37 (m, 4H), 0.88 (t, J =

    6.2 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ 162.56 (d, J = 248.5 Hz), 143.59 –

    143.81 (m), 132.39 (dd, J = 33.1, 7.8 Hz), 123.30 (dd, J = 272.4, 2.7 Hz), 120.52

    – 120.74 (m), 118.50 (d, J = 21.6 Hz), 111.39 (dq, J = 24.5, 3.8 Hz), 52.62,

    49.08, 29.35, 29.23, 22.48, 13.92. HRMS (APCI+, m/z): calculated for C13H18F4N

    [M+H]+: 264.13699; found: 264.13668.

    3-Chloro-4-fluoro-N-pentylbenzylamine (8fa): Synthesized according to

    General procedure A. 3-Chloro-4-fluorobenzylamine (0.069 ml, 0.5 mmol) affords

    8fa (0.050 g, 43% yield). Yellow oil obtained after column chromatography (SiO2,

    n-pentane/EtOAc 50:50 to 0:100). 1H NMR (400 MHz, CDCl3) δ 7.33 – 7.43 (m,

  • Alkylation of amines with alcohols

    43

    1H), 7.14 – 7.22 (m, 1H), 7.02 – 7.11 (m, 1H), 3.73 (s,

    2H), 2.59 (t, J = 7.5 Hz, 2H), 1.93 – 2.08 (br.s, 1H), 1.43

    – 1.58 (m, 2H), 1.25 – 1.37 (m, 4H), 0.83 – 0.93 (m, 3H). 13C NMR (100 MHz, CDCl3) δ 157.07 (d, J = 247.5 Hz),

    137.34, 130.15, 127.69 (d, J = 7.0 Hz), 120.69 (d, J = 17.8

    Hz), 116.29 (d, J = 20.9 Hz), 52.71, 49.26, 29.56, 29.45,

    22.54, 13.99.

    1-(3-fluorobenzyl)pyrrolidine (9cm): Synthesized according to

    General procedure A. 3-Fluorobenzylamine (0.057 ml, 0.5 mmol)

    affords 9cm (0.054 g, 60% yield). Yellow oil obtained after column

    chromatography (Al2O3, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CD2Cl2) δ 7.21 – 7.32 (m, 1H), 7.02 – 7.15 (m, 2H), 6.93

    (td, J = 8.6, 2.2 Hz, 1H), 3.59 (s, 2H), 2.32 – 2.60 (m, 4H), 1.67 –

    1.82 (m, 4H). 13C NMR (100 MHz, CD2Cl2) δ 162.87 (d, J = 245.1 Hz), 142.06 (d,

    J = 6.2 Hz), 129.56 (d, J = 8.4 Hz), 124.27 (d, J = 2.8 Hz), 115.57 (d, J = 21.4

    Hz), 113.70 (d, J = 21.4 Hz), 60.14, 54.14, 23.45. The physical data were identical

    in all respects to those previously reported.[45]

    1-(3-fluorobenzyl)piperidine (9cl): Synthesized according to

    General procedure A. 3-Fluorobenzylamine (0.057 ml, 0.5 mmol)

    affords 9cl (0.060 g, 62% yield). Yellow oil obtained after column

    chromatography (Al2O3, n-pentane/EtOAc 100:0 to 95:5). 1H NMR (400

    MHz, CD2Cl2) δ 7.25 – 7.37 (m, 1H), 7.07 – 7.25 (m, 2H), 6.90 – 7.07

    (m, 1H), 3.60 (s, 2H), 2.30 – 2.72 (m, 4H), 1.56 – 1.77 (m, 4H), 1.39

    – 1.55 (m, 2H). 13C NMR (100 MHz, CD2Cl2) δ 163.34 (d, J = 244.1 Hz),

    142.76 (d, J = 7.0 Hz), 129.87 (d, J = 8.3 Hz), 124.86 (d, J = 2.7 Hz), 115.81 (d,

    J = 21.2 Hz), 113.80 (d, J = 21.2 Hz), 63.48 (d, J = 1.9 Hz), 54.93, 53.84, 26.51,

    24.82. The physical data were identical in all respects to those previously

    reported.[46]

    1-(3-fluorobenzyl)azepane (9ck): Synthesized according to

    General procedure A. 3-Fluorobenzylamine (0.057 ml, 0.5 mmol)

    affords 9ck (0.069 g, 67% yield). Yellow oil obtained after column

    chromatography (Al2O3, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 7.18 – 7.30 (m, 1H), 7.00 – 7.18 (m, 2H), 6.84

    – 6.98 (m, 1H), 3.63 (s, 2H), 2.45 – 2.72 (m, 4H), 1.43 – 1.84 (m,

    8H). 13C NMR (100 MHz, CDCl3) δ 162.94 (d, J = 244.9 Hz), 143.01 (d, J = 6.0

    Hz), 129.40 (d, J = 8.1 Hz), 124.06 (d, J = 2.7 Hz), 115.31 (d, J = 21.2 Hz),

    113.49 (d, J = 21.3 Hz), 62.24 (d, J = 1.9 Hz), 55.61, 28.26, 26.97. HRMS (APCI+,

    m/z): calculated for C12H19FN [M+H]+: 208.14960; found: 208.14928.

    1-(3-chlorobenzyl)azepane (9bk) : Synthesized according to

    General procedure A. 3-Chlorobenzylamine (0.061 ml, 0.5 mmol)

    affords 9bk (0.095 g, 85% yield). Yellow oil obtained after column

    chromatography (Al2O3, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 7.36 (s, 1H), 7.14 – 7.25 (m, 3H), 3.61 (s, 2H),

    2.55 – 2.66 (m, 4H), 1.58 – 1.70 (m, 8H). 13C NMR (100 MHz, CDCl3)

  • Chapter 2

    44

    δ 142.36, 134.01, 129.32, 128.66, 126.82, 126.72, 62.15, 55.58, 28.20, 26.98.

    HRMS (APCI+, m/z): calculated for C13H19ClN [M+H]+: 224.12005; found:

    224.11979.

    1-(3-(trifluoromethyl)benzyl)azepane (9dk): Synthesized

    according to General procedure A. 3-Trifluoromethylbenzylamine

    (0.072 ml, 0.5 mmol) affords 9dk (0.087 g, 68% yield). Yellow oil

    obtained after column chromatography (Al2O3, n-pentane/EtOAc

    100:0 to 95:5). 1H NMR (400 MHz, CDCl3) δ 7.26 (s, 1H), 7.55 (d, J

    = 7.5 Hz, 1H), 7.49 (d, J = 7.8 Hz, 1H), 7.41 (m, 1H), 3.69 (s, 2H),

    2.51 – 2.70 (m, 4H), 1.53 – 1.72 (m, 8H). 13C NMR (100 MHz, CDCl3) δ 143.31,

    131.91, 130.43 (q, J = 21.7 Hz), 128.48, 125.25 (q, J = 3.8 Hz), 124.33 (d, J =

    272.4 Hz), 123.53 (q, J = 3.8 Hz), 62.22, 55.60, 28.27, 26.98. HRMS (APCI+,

    m/z): calculated for C14H19F3N [M+H]+: 258.14641; found: 258.14615.

    1-(3-fluoro-5-(trifluoromethyl)benzyl)azepane (9ek): Synthesized

    according to General procedure A. 3-Fluoro-5-

    trifluoromethylbenzylamine (0.073 ml, 0.5 mmol) affords 9ek

    (0.091 g, 66% yield). Yellow oil obtained after column

    chromatography (Al2O3, n-pentane/EtOAc 100:0 to 95:5). 1H NMR

    (400 MHz, CDCl3) δ 7.40 (s, 1H), 7.32 (d, J = 9.7 Hz, 1H), 7.18 (d,

    J = 8.2 Hz, 1H), 3.67 (s, 2H), 2.53 – 2.69 (m, 4H), 1.52 – 1.73

    (m, 8H). 13C NMR (100 MHz, CDCl3) δ 162.58 (d, J = 247.7), 144.78 (d, J = 7.1

    Hz), 132.08 (dd, J = 33.0, 8.3 Hz), 123.46 (dd, J = 272.5, 3.1 Hz), 120.67 –

    120.86 (m), 118.65 (d, J = 21.2 Hz), 111.04 (dq, J = 24.9, 3.8 Hz), 61.90 (d, J =

    1.6 Hz), 55.68, 28.36, 26.95. HRMS (APCI+, m/z): calculated for C14H18F4N

    [M+H]+: 267.13699; found: 276.13666.

    Piribedil (12): Synthesized according to General

    procedure B. 1-(2-Pyrimidyl)piperazine (10) (0.082 g,

    0.5 mmol) and 4 eq. piperonyl alcohol (11) affords 12

    (0.080 g, 54% yield). Yellow oil obtained after column

    chromatography (Al2O3, n-pentane/EtOAc 90:10 to

    50:50). 1H NMR (400 MHz, CDCl3) δ 8.28 (d, J = 4.7Hz, 2H), 6.88 (s, 1H), 6.75 (s,

    2H), 6.45 (t, J = 4.7Hz, 1H), 5.93 (s, 2H), 3.81 (t, J = 4.9 Hz, 4H), 3.44 (s, 2H),

    2.47 (t, J = 5.0Hz, 4H). 13C NMR (100 MHz, CDCl3) δ 161.60, 157.62, 147.62,

    146.62, 131.74, 122.19, 109.66, 109.45, 107.83, 100.85, 62.80, 52.77, 43.62.

    The physical data were identical in all respects to those previously reported.[47,48]

  • Alkylation of amines with alcohols

    45

    Figure 7 NOESY-NMR spectrum of isolated compound 3qa.

  • Chapter 2

    46

    References

    [1] G. Guillena, D. J. Ramon, M. Yus, Chem. Rev., 2010, 110, 1611–1641.

    [2] a) A. Hager, N. Vrielink, D. Hager, J. Lefranc, D. Trauner, Nat. Prod. Rep., 2016, 33,

    491–522; b) J. L. McGuire, Pharmaceuticals: Classes, Therapeutic agents, areas of

    Application, Vol. 1–4, Wiley-VCH, 2000.

    [3] S. Bähn, S. Imm, L.Neubert, M. Zhang, H. Neumann, M. Beller, ChemCatChem, 2011,

    3, 1853–1864.

    [4] A. J. A. Watson, J. M. J. Williams, Science, 2010, 329, 635–636.

    [5] C. Gunanathan, Y. Ben-David, D. Milstein, Science, 2007, 317, 790-792.

    [6] P. J. Deuss, K. Barta, J. G. de Vries, Catal. Sci. Technol., 2014, 4, 1174-1196.

    [7] K. Barta, P. C. Ford, Acc. Chem. Res., 2014, 47, 1503–1512.

    [8] B. M. Trost, Science, 1991, 254, 1471–1477.

    [9] P. T. Anastas, J. C. Warner, Green Chemistry: Theory and Practice, Oxford University

    Press, 1998.

    [10] R. Grigg, T. R. B. Mitchell, S. Sutthivaiyakit, N. J. Tongpenyai, J. Chem. Soc. Chem.

    Commun., 1981, 611.

    [11] M. H. S. A. Hamid, C. L. Allen, G. W. Lamb, A. C. Maxwell, H. C. Maytum, A. J. A.

    Watson, J. M. J. Williams, J. Am. Chem. Soc., 2009, 131, 1766–1774.

    [12] S-I. Murahashi, K. Kondo, T. Hakata, Tetrahedron Lett., 1982, 23, 229–232.

    [13] K.-I. Fujita, T. Fujii, R. Yamaguchi, Org. Lett., 2004, 6, 3525–3528.

    [14] M. H. S. A. Hamid, P. A. Slatford, J. M. J. Williams, Adv. Synth. Catal., 2007, 349,

    1555–1575.

    [15] C. Gunanathan, D. Milstein, Science, 2013, 341, 1229712.

    [16] G. E. Dobereiner, R. H. Crabtree, Chem. Rev., 2010, 110, 681–703.

    [17] R. Martinez, D. J. Ramon, M. Yus, Org. Biomol. Chem., 2009, 7, 2176–2181.

    [18] S. Enthaler, K. Junge, M. Beller, Angew. Chem. Int. Ed., 2008, 47, 3317–3321.

    [19] M. Bala, P. K. Verma, U. Sharma, N. Kumar, B. Singh, Green Chem., 2013, 15,

    1687–1693.

    [20] Y. Zhao, S. W. Food, S. Saito, Angew. Chem. Int. Ed., 2011, 50, 3006–3009.

    [21] H.-J. Knölker, E. Baum, H. Goesmann, R. Klauss, Angew. Chem. Int. Ed., 1999, 38,

    2064-2066.

    [22] C. P. Casey, H. Guan, J. Am. Chem. Soc., 2007, 129, 5816-5817.

    [23] A. Pagnoux-Ozherelyeva, N. Pannetier, M. D. Mbaye, S. Gaillard, J.-L. Renaud, Angew.

    Chem. Int. Ed., 2012, 51, 4976-4980.

    [24] a) T. N. Plank, J. L. Drake, D. K. Kim, T. W. Funk, Adv. Synth. Catal., 2012, 354,

    597-601; b) M. G. Coleman, A. N. Brown, B. A. Bolton, H. Guan, Adv. Synth. Catal.,

    2010, 352, 967–970.

    [25] a) S. Zhou, S. Fleischer, K. Junge, M. Beller, Angew. Chem. Int. Ed., 2011, 50, 5120–

    5124; b) A. Quintard, T. Constantieux, J. Rodriguez, Angew. Chem. Int. Ed., 2013,

    52, 12883–12887.

    [26] a) 1H NMR spectrum of commercially available pentane-1-al in toluene-d8 at 100 °C.

    The chemical shift of the indicative aldehyde proton (HA) is at 9.40 ppm; b) 1H NMR

    spectrum of imine 4aa, in situ generated from 0.12 mmol p-methoxy-aniline (1a)

    and 0.50 mmol pentane-1-al. The compounds were solubilized in 0.6 ml toluene-d8

    at room temperature, after 0.5 hour the 1H NMR spectrum was recorded at 100 °C.

    The indicative aldehyde proton HA is found at 9.40 ppm. The indicative imine proton

    HI is found at 7.67 ppm, in agreement with the literature: J. C. Anderson, G. P. Howell,

    R. M. Lawrence, C. S. Wilson, J. Org. Chem., 2005, 70, 5665–5670 (7.63 ppm for

    4-methoxy-N-hexylaniline in C6D6 at r.t.).

    [27] C. P. Casey, H. Guan, J. Am. Chem. Soc., 2009, 131, 2499–2507.

    [28] K. Watanabe, N. Yamagiwa, Y. Torisawa, Org. Process Res. Dev., 2007, 11, 251–

    258.

    [29] M. A. Berliner, S. P. A. Dubant, T. Makowski, K. Ng, B. Sitter, C. Wager, Y. Zhang,

    Org. Process Res. Dev., 2011, 15, 1052–1062.

    [30] Although the benzene ring is not conjugated to the imine, the effect from the

    substitutions to the reduction rate of imine might be non-linear.

  • Alkylation of amines with alcohols

    47

    [31] M. Jaber, S. W. Robinson, C. Missale, M. G. Caron, Neuropharmacology, 1996, 35,

    1503–1519.

    [32] T. N. Plank, J. L. Drake, D. K. Kim, T. W. Funk, Adv. Synth. Catal., 2012, 354, 597-

    601.

    [33] D. Jiang, H. Fu, Y. Jiang, Y. Zhao, J. Org. Chem., 2007, 72, 672-674.

    [34] E. J. Debeer, J. S. Buck, W. S. Ide, A. M. Hjort, J. Pharm. Exp. Ther., 1936, 57, 19-

    33.

    [35] J. Barluenga, A. M. Bayón, G. Asensio, J. Chem. Soc. Chem. Commun., 1983, 1109-

    1110.

    [36] J. Lau, J. T. Kodra, M. Guzel, K. C. Santosh, A. M. M. Mjalli, R. C. Andrews, D. R.

    Polisetti, PCT Int. Appl. WO 2003055482 A1, 2003.

    [37] J. Wu, C. Wang, W. Tang, A. Pettman, J. Xiao, Chem. Eur. J., 2012, 18, 9525-9529.

    [38] Q. Shen, S. Shekhar, J. P. Stambuli, J. F. Hartwig, Angew. Chem. Int. Ed., 2005, 44,

    1371-1375.

    [39] O. Saidi, A. J. Blacker, M. M. Farah, S. P. Marsden, J. M. J. Williams, Angew. Chem.

    Int. Ed., 2009, 48, 7375-7378.

    [40] J. C.-H. Yim, J. A. Bexrud, R. O. Ayinla, D. C. Leitch, L. L. Schafer, J. Org. Chem.,

    2014, 79, 2015-2028.

    [41] M. Yang, F. Liu, J. Org. Chem., 2007, 72, 8969-8971.

    [42] M. Utsunomiya, J. F. Hartwig, J. Am. Chem. Soc., 2004, 126, 2702-2703.

    [43] S. Lu, J. Wang, X. Cao, X. Li, H. Gu, Chem. Comm., 2014, 50, 3512-3515.

    [44] W. R. Meindl, E. V. Angerer, H. Schoenenberger, G. Ruckdeschel, J. Med. Chem.,

    1984, 27, 1111-1118.

    [45] S. D. Banister, L. M. Rendina, M. Kassiou, Bioorg. Med. Chem. Lett., 2012, 22, 4059-

    4063.

    [46] L. G. Erwan, D. Alexandre, M. Thierry, T. Michel, Synthesis, 2010, 249-254.

    [47] M. Haniti, S. A. Hamid, J. M. J. Williams, Tetrahedron Lett., 2007, 48, 8263–8265.

    [48] X. Cui, X. Dai, Y. Deng, F. Shi, Chem. Eur. J., 2013, 19, 3665-3675.

  • Chapter 2

    48

    Chapter 2