smc044

Embed Size (px)

Citation preview

  • 8/15/2019 smc044

    1/13

    Scanning Electron Microscopy

    Zhan-Ting Li and Xin ZhaoChinese Academy of Sciences, Shanghai, China

    1 Introduction 1

    2 Supramolecular Gels 23 Supramolecular Liquid Crystals 5

    4 Supramolecular Polymers 6

    5 Vesicles 7

    6 Nanotubes, -Pores, -Ribbons, -Rods, and -Spheres 8

    7 Supramolecular Chirality 11

    8 Conclusion 12

    References 13

    1 INTRODUCTION

    The scanning electron microscope (SEM) is an electron

    microscope that uses electron beams to scan a sample and

    form an image of the sample surface.1–3 The electrons

    interact with the atoms of the sample to produce various

    signals, which include secondary electrons, primary back-

    scattered electrons, characteristic X-rays, Auger electrons,

    cathodoluminescence, and specimen current and transmitted

    electrons, providing information about the sample’s sur-

    face topography, composition, and electrical conductivity.

    Although all these signals are present in SEM, only the

    first three signals, especially the secondary electrons, are

    commonly used in commercial instruments, which can pro-

    duce very high-resolution images of a sample surface. The

    first SEM image was obtained by Knoll in 1935,4 while

    the first commercial instrument was developed in 1965 by

    the Cambridge Instrument Company, named “Stereoscan.”

    Since then, it has found wide applications in the biological,

    materials, and chemical researches.5

    There are many advantages of using SEM in the detectionof a sample: (i) It is relatively cheap and widely available.

    (ii) Owing to the very narrow electron beam, it has a

    large depth of field, which allows a large amount of 

    the sample to be in focus at one time and yields a

    characteristic three-dimensional appearance. (iii) SEM can

    produce images of very high resolution. Greater than

    500 000 times magnification can be achieved, which means

    that several nanometers in size can be revealed. (iv) The

    magnification rate of SEM can be easily set successively.

    Therefore, for a sample, a low magnification can be used

    to obtain the whole picture, while the high magnification

    may be set to observe the detailed structures. (v) SEM has

    a large depth of focus, which yields a characteristic three-

    dimensional image that is useful in obtaining information

    on the surface structure of a sample.

    Since the secondary electrons are of low energy, their

    trajectories can be easily affected by electromagnetic fields.

    As a result, a charge buildup on the surface of the sample

    can change the path of the secondary electrons. To avoid

    this, the surface of the sample must be conductive. Thus,

    the electroconductive sample can be measured directly. For

    non-electroconductive samples, shadowing methods have

    been developed to coat the samples with a thin layer

    of metal. Thus, the preparation of the sample for SEM

    measurement is, in general, simple.

    The conventional SEM measurement carried out in a

    vacuum has disadvantages and may also result in artifacts.Moreover, the coating can obscure the fine surface structure

    details of some non-electroconductive samples. In the

    1980s, the environmental scanning electron microscope

    (ESEM) was developed,6 which permits the imaging of 

    wet systems with no prior specimen preparation. Since the

    sample environment can be dynamically altered, hydration

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©  2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    2/13

    2   Techniques

    and dehydration processes can be followed as they happen

    in the sample chamber. Also in the 1980s, cryo (cold)

    SEM became commercially available,7 allowing samples

    to be preserved and recorded at below ambient temperature

    (typically between  −100 and  −175◦

    C). Another improved

    technique, the field emission (FE) SEM can produce clearer,

    less electrostatically distorted images with spatial resolutiondown to 1.5nm.3

    Supramolecular chemistry focuses on the chemistry

    beyond the molecules and the bottom-up building of 

    ordered assembled systems from molecular subunits or

    components. The binding process of a synthetic receptor

    for a single molecular or ionic species is difficult to observe

    directly by any electron microscopic techniques. However,

    an ocean of assembled systems exhibit ordered morpholo-

    gies on the nano- to microscales. With more and more inter-

    est focusing on the complicated systems based on molecular

    recognition and self-assembly, the structural information

    within this range of scales is becoming increasingly impor-

    tant. SEM has become a powerful and routine analyticaltechnique for collecting such information. In most studies,

    SEM is used together with other microscopic techniques

    such as AFM, TEM, and STM, which guarantees a com-

    plete understanding of the morphology of the sample.

    2 SUPRAMOLECULAR GELS

    Gels are a colloidal phase state in which a small amount of 

    a solid-like network immobilizes the bulk flow of a larger

    amount of liquid phase, which may be considered as a spe-

    cial supramolecular system. Both polymeric and molecular

    building blocks can form the three-dimensional networks.

    Typically, the networks consist of fibrillar structures. For

    molecular gels, these fibers are formed by the self-assembly

    of molecular blocks by complementary noncovalent interac-

    tions. The formation of the gels is usually judged simply by

    visual observations (vial inversion test). However, their fine

    structures can be investigated only by using various micro-

    scopes. The size of the fibers of gels generally ranges from

    several nanometers to a few tens of micrometers, which is

    well covered by SEM. Therefore, SEM is widely used for

    observing the morphology of the fibers.

    An early example concerns organogels of compound

    1   with aliphatic solvents, reported by Terech   et al.8

    The small-angle X-ray and neutron scattering techniques

    demonstrated that the gel networks resulted from the entan-glement of long, solid-like rigid fibers, while SEM was used

    to characterize the morphologies of the xerogels (Figure 1).

    Owing to collapses of the brittle structures in the 3D net-

    work during the shrinking step caused by evaporation of 

    the liquid, SEM focused on the general shapes and mor-

    phologies rather than on the absolute quantities such as the

    OH

    H   1

    10 µm200 µm

    20 µm

    10 µm

    2 µm

    20 µm

    (a)

    (b)

    (c)

    (d)

    (e)

    (f)

    Figure 1   SEM images of fractured xerogels of   1   in cyclo-hexane: (a) random repartition of very long and rigid fibers;(b) detail showing the presence of thinner and more flexiblefibrils; (c) collapsed network of the phase-separated solid of 1 /heptane gel, showing fibers emanating from a central point;(d) thick bundles of fibers; (e and f) high orientation degree of fibers. (Reproduced from Ref. 8.  ©   American Chemical Society,1998.)

    diameters, lengths, or topologies of the fibers. Figure 1(a)

    shows the formation of very long and rigid fibers of variable

    thickness (0.1– 0.2µm) which are entangled in a porous

    matrix. Figure 1(b) reveals that thinner, more flexible fib-

    rils (0.05µm) are also present, while Figure 1(c) exhibits

    a special morphology where fibers are emanating from a

    central point. When stored for a long time, the gel can

    evolve toward a solid–liquid-phase separation, which is

    also confirmed by SEM (Figure 1d–f), because the micro-

    graphs show that the phase-separated solid is made of fibers

    associated in locally aligned bunches. This observation also

    confirms that the true equilibrium state of such very long

    and rigid fibers is the separated biphasic system, which

    has a higher degree of orientation than the xerogel throughthe formation of oriented bundles by collapses of separated

    fibers of the xerogel.

    The magnification rate of SEM can be set successively.

    This is not necessary for observing the morphology of 

    supramolecular gels. However, it does provide access to

    recording images at different magnifications. For example,

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©   2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    3/13

    Scanning electron microscopy   3

    N

    N

    O O

    OO

    OCH3O17

    N

    EtEt

    Et Et

    N

    N

    OO

    O O

    O

    17

    N

    Et Et

    EtEt

    2

    OCH3

    0.5 µm

    5 µm

    200 nm

    2 µm

    (a) (b)

    (d)(c)

    Figure 2   Cryo-SEM images of the gel of  2  (8 mM, water/THF

    mixture (80: 20)) at different magnifications. (a) Nanoporousstructure; inset: vial inversion test. (b) Image at higher magni-fication showing the three-dimensional network of nanofibers.The smallest fibers are about 6.1 nm in diameter (yellow arrow).(c) Whirls with diameters of 10–15 µm. (d) Directional arrange-ment of fibers within a “microstream” in the gel. (Reproducedfrom Ref. 9.  ©   American Chemical Society, 2009.)

    Rybtchinski reported that compound   2   gelated water/THF

    (tetrahydrofuran) mixtures due to the hydrophobically

    driven aromatic stacking.9 Cryo-SEM images revealed an

    interconnected porous structure of the gel (Figure 2a),

    in which nanofibers created a three-dimensional network 

    (Figure 2b). It also showed that the smallest fibers had

    widths of about 6.1 nm, and their actual width (subtract-ing the thickness of the metal layer used for gel imag-

    ing   −0.6 nm) was about 5.5 nm, which is similar to the

    size observed in the case of the solution-phase network.

    Furthermore, thicker fibers of various diameters were also

    observed. The images at lower magnifications revealed the

    presence of whirls and streams (Figure 2c and d). These

    anisotropic regions are several micrometers large, and

    demonstrate a certain long-range order of the supramolecu-

    lar fibers within the gel. On the basis of these observations

    and also molecular modeling, the authors proposed a hier-

    archically assembling mechanism for the gelation process.

    At the first level, the hydrophobic and stacking interac-

    tions caused face-to-face stacking of compound  2  into smallaggregates of 8–10 molecules. The stacked molecules were

    shifted with respect to each other, due to the steric bulk of 

    the aliphatic chains. At the second level, the hydropho-

    bic effect was the driving force for further aggregation,

    which was driven by the aliphatic side chains that, as a

    result of aromatic stacking, formed a substantial hydropho-

    bic domain. Their interaction resulted in fibers with distinct

    segmentation. Then, the fibers assembled into entangled

    bundles, while branching out of these bundle accounted for

    the creation of junctions.

    The superfine structures of organogels can also be stud-

    ied by using SEM. For example, Zhang   et al. found

    that compound   3   gelated cyclohexane through the for-mation of an entangled network of thin solid fibers with

    diameters of about 40–80 nm and lengths up to tens

    of microns (Figure 3a).10 The structure of the gel was

    changed when 7,7,8,8-tetracyanoquinodimethane (TCNQ)

    was added because it formed a charge-transfer complex

    with the tetrathiafulvalene unit of   3. Subsequently, the

    entangled thin fibers of the gel was transformed into a tube-

    like structure with diameters of about 20 –60 nm, which was

    clearly shown by using SEM (Figure 3b).

    For rigid and partially rigid gelators, their fibrous struc-

    tures can be readily observed by using SEM. However,

    SEM focuses on the three-dimensional surface information.

    For proposing a rational assembling pattern, in many stud-

    ies, other techniques have to be used.11,12 Lu  et al. have

    reported that compound  4  gelated hydrocarbons (Figure 4).

    SEM demonstrated that the molecules in the gel phase

    were self-assembled into one-dimensional nanofibers with a

    25– 100 nm width, which further cross-linked to form three-

    dimensional networks (Figure 4a).11 The small-angle X-ray

    diffraction of the xerogel illustrated that the molecules

    were packed into the lamellar structure (Figure 4b), which,

    together with semiempirical quantum calculation, supported

    that the molecules adopted a one-dimensional molecular

    packing pattern to self-assemble into thin fibers (Figure 4d),

    which were further wound or laced to give the wide

    nanofibers (Figure 4a).

    Most of the reported gelators are single molecules. Two-or multicomponent systems can also gelate liquids if they

    are able to assemble three-dimensional networks through

    the formation of similar fibrous structures, which can also

    be characterized by using SEM.13,14 For example, Smith

    et al. found that both compound   5   and its mixture with

    diamine   6   gelated toluene.13 The SEM images of their

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©  2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    4/13

    4   Techniques

    S

    S

    S

    SS

    S

    SNH

    NH

    O

    3

    (a) (b)

    NONE SEI 300 V WD 6.7 nm1 µm× 10.000NONE SEI 300 V WD 6.7 nm1 µm× 4.10

    1 µm

    100nm

    Figure 3   SEM image of (a) the cyclohexane gel of  3  and (b) that of the charge-transfer complex of  3  with TCNQ. The inset clearlyshows the tube-like superfine structures of the complex gel. (Reproduced from Ref. 10.  ©   American Chemical Society, 2005.)

    (a)

    (b)

    2 4 6 8

    Cu-KR (2q  / °)

    3.01 nm

    1.49 nm

    1 µm

    (c)

    (d)

    (e)

    3.01 nm

       I  n   t  e  n  s   i   t  y   (  a .  u .

       )

    O

    OO

    O OO

    O

    O

    HOOH

    N N

    4

    Figure 4   (a) SEM image and (b) SAXD pattern of xerogel 4 obtained from cyclohexane. (c) Molecular structure of  4. (d) Unimolecularstacking with a length of 3.01 nm in the gel. (e) Proposed molecular packing model along the growing direction of the gel fiber.(Reproduced from Ref. 11.  ©   American Chemical Society, 2009.)

    xerogels revealed that both samples formed fibrous struc-tures (Figure 5) and the addition of the second compo-

    nent dramatically changed the nanoscale morphology of 

    the assembled superstructure. On the basis of the observa-

    tions, they proposed that the formation of the complex made

    the network more interpenetrate and the nanoscale fibers

    narrower because of the presence of more “connecting

    points,” that is, the acid-amino electrostatic interactions of their complex.

    Gels usually remain in their phase state for some time

    and eventually collapse to amorphous precipitates. Tang

    et al. found that the hydrogel of the 1 : 2 mixture of   7

    and   8   could spontaneously transfer into macroscopic crys-

    tals during storage.15 One related issue was to observe

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©   2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    5/13

    Scanning electron microscopy   5

    NH

    O

    NH

    C11H23

    C11H23

    O

    HO

    O

    NH(CH2)12NH2NH

    C11H23

    O

    NH

    C11H23

    O

    OH

    O5 6 5

    (a) (b)

    Figure 5   SEM images of   5   in the absence ((a) scale bar =20µm) or presence ((b) scale bar = 2µm) of  6. (Reproduced fromRef. 13.  ©  Royal Society of Chemistry, 2005.)

    O

    O

    O

    O

    O

    O

    O

    O

    H

    H

    H

    H

    N

    N

    OH

    OH7

    8

    8

    100 µm100 µm

    100 µm 100 µm

    50 µm 50 µm

    (a)

    (b)

    (c)

    a'

    b'

    c'

    Crystal

    Fiber

    Figure 6   ESEM images of hydrated samples after storing thegel (formed at 5 wt%,   7 : 8   =   1:2) for (a) 1h, (b) 12h, and(c) 36 h. a, b, and c represent the dehydrated samples after 1 h,12 h, and 36 h storage, respectively. (Reproduced from Ref. 15.©   American Chemical Society, 2008.)

    the transferring process from the gels to the crystals. As

    it is difficult to use conventional SEM to obtain high-

    resolution images of naturally hydrated gels under high

    vacuum, ESEM was applied to study the transition. The

    samples were observed by using ESEM at constant vapor

    pressures of 560 Pa (for hydrated samples) and 410 Pa (for

    dehydrated samples) (Figure 6). On storing for 1 h, the sam-

    ple displayed an entangled fibrillar network filled with water

    (Figure 6a), which is typical of ordinary gels. After 12 h,

    fibers and crystals were present simultaneously in the sam-

    ple (Figure 6b). After 36 h, only prismatic crystals remained

    (Figure 6c). The transition was clearer for dehydrated sam-

    ples (Figure 6a–c ) and could also be observed using SEM,

    but only for the xerogel.

    3 SUPRAMOLECULAR LIQUID

    CRYSTALS

    Supramolecular liquid crystals are a state of matter that

    consists of two or more components and exhibit prop-

    erties between those of a conventional liquid and those

    of a solid crystal. The molecular components bind each

    other to form a liquid crystalline mesophase, which is

    typically characterized by polarized optical microscopy

    (POM), X-ray diffraction (XRD), and differential scan-

    ning calorimetry (DSC). SEM is useful for studying their

    assembling structures on the surface and helps establish

    their assembling patterns. For example, Yagai  et al. have

    reported that POM, XRD, and DSC experiments on the

    films revealed that complexes of   9  and   10a   and  b  formed

    identical fan-shape textures characteristic of columnar liq-

    uid crystalline phases.16 The complexes might form a 1 : 1

    Hamilton-type complex or extended supramolecular poly-

    mers. Dynamic light scattering (DLS) provided evidence

    that the Hamilton-type complexes were formed in mil-

    limolar concentrations. Optical microscopy (OM), POM,

    and SEM images were further obtained for their solu-

    tions in cyclohexane (Figure 7). OM and POM images

    (Figure 7a–d) showed that the fibers that were in parallel

    to either the polarizer or the analyzer (those indicated by

    arrows) were not visible under crossed-polarizer condition,

    indicating the uniaxial nature of the fibers. The FE-SEM

    image illustrated that the fibers were composed of bundled

    thinner fibers with diameters less than 100 nm (Figure 7e),

    and the cross-sectional image showed that the narrowerfibers possessed a solid interior (Figure 7f). Because the

    fibers gave XRD peaks almost identical to those of their

    films, the SEM results indicated that the fibers consisted of 

    hexagonally packed columns of disk-like 1 : 1 complexes,

    and thus supported the formation of Hamilton-type binding

    pattern.

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©  2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    6/13

    6   Techniques

    NR2

    N N

    OO

    N

    N

    N

    N

    RO

    R2N N

    H

    H

    HO

    N

    N

    N

    N

    OR

    NR2N

    H

    H

    H

    (i) Hamilton-type complexes

    R2N

    N N

    OO

    N

    N

    N

    N

    R2N N

    H

    Ar

    H

    H

    O

    N

    N

    N

    N

    NR2N

    H

    Ar

    H

    H

    R2N

    N N

    OO

    N

    N

    N

    N

    R2N N

    H

    Ar

    H

    HO

    N

    N

    N

    N

    NR2N

    H

    Ar

    H

    H

    (ii) Extended supramolecular polymers

    910

    10

    9

    9

    8

    10

    R2N

    HN NH

    OO

    O

    9

    NH

    N

    NNRO

    NR2

    NH

    (CH2)n 

    HNN

    N N OR

    NR2

    HN

    10a: n  = 5   10b: n  = 7

    R = n -C12H25

    100 µm

    100 µm

    5 µm 1 µm

    100 µm

    A

    P

    A

    P

    (a)

    (c)

    (e)

    (b)

    (d)

    (f)

    100 mm

    Figure 7   OM (a, b) and POM images (c, d) of nanofibers of 9·10a  (a, c) and  9·10b  (b, d) grown from cyclohexane (5.0mM),and SEM images of nanofibers of  9·10b (e, f). (Reproduced fromRef. 16.  ©   American Chemical Society, 2007.)

    The difference in the thin film morphology exhibited

    on the SEM images can also be used to reveal the effect

    of the molecular structures on the self-assembling prop-

    erties. Asha   et al. have reported that compound   11a– d

    produced a liquid crystalline phase.17 SEM images were

    used to reveal the effect of the functional group ester ver-

    sus amide and the flexible alkyl terminal chains on theself-assembling properties (Figure 8). The image of   11a

    (Figure 8a) showed one-dimensional rod-like supramolecu-

    lar structures with diameters of 0.5–1.0µm and lengths of 

    several micrometers. The rods tended to aggregate into huge

    rods, as shown in the inset. The image of  11b  (Figure 8b),

    which showed the maximum aggregation tendency in the

    solution, gave supramolecular organization of rods stacked

    several micrometers long. In regions where isolated rods

    were identified, the thinnest rods had widths of about

    150nm and lengths of a few tens of micrometers, lead-

    ing to aspect ratios (length over width) of magnitude 100.

    This long aspect ratio compared to that of the unsubsti-

    tuted   11a   was attributed to the longer molecular lengthof   11b   with respect to   11a. The SEM of   11c   (Figure 8c)

    showed one-dimensional ball-like supramolecular structures

    of 3–4µm diameter that were interconnected by nanometer-

    scale rods (indicated by arrows), which was ascribed to

    the bulky tridodecyl substitution at the terminal benzene

    rings. The image of ester derivative   11d   (Figure 8d) dis-

    played a leaf-like pattern, 1.8–3 µm in diameter and sev-

    eral micrometers in length, which is distinctly different

    from the rod- or ball-like aggregates formed by the amide

    series. It was proposed that the absence of an additional

    building force such as the hydrogen bonding in the amide

    series resulted in such a different morphology, because

    11d   had only   π –π   interaction to aid in its self-assembly

    process.

    4 SUPRAMOLECULAR POLYMERS

    A supramolecular polymer is a polymer whose monomeric

    repeating units are held together by noncovalent bonds. For

    main-chain supramolecular polymers, the linear backbones

    may form ordered structures. If the size of these assem-

    bled architectures falls into the range covered by SEM,

    then SEM can be used to investigate their morphology on

    the surface, which provides useful information for estab-

    lishing the assembling patterns. Haino  et al. reported that

    compounds   12   and   13   formed supramolecular polymersin solution through the calix[5]arene– C60   stacking inter-

    actions.18 The size and morphology of the composite on

    the surface were confirmed by the SEM image of their 1 : 1

    solution (Figure 9a and b). The thicker entwined fibers had

    lengths of more than 100µm and widths of 250–500 nm

    (Figure 9b), indicating that ditopic host  13 iteratively bound

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©   2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    7/13

    Scanning electron microscopy   7

    N N

    O

    O

    O

    O

    X X

    N

    H

    O

    11a: X = N

    H

    O

    11b: X =

    NH

    O

    11c: X =

    OC12H25

    OC12H25

    OC12H25

    OC12H25

    OC12H25 O

    O

    11d: X =

    (a) (b)

    (c) (d)

    7 kV × 5,000 5 µm 2690 RRLSEM 15 kV × 6,000 2 µm 2670 RRLSEM

    10 kV 10 kV× 5,000 5 µm 2680 RRLSEM × 5,000 5 µm 2683 RRLSEM

    Figure 8   Scanning electron micrograph (SEM) images of self-assembled (a) 11a , (b)  11b, (c)  11c , and (d)  11d  drop-cast from toluene.

    (Reproduced from Ref. 17. ©   American Chemical Society, 2008.)

    to dumbbell fullerene   12   to create a two-dimensional

    nanonetwork.

    Yagai   et al. have reported that compounds   14   and   15

    formed supramolecular polymers through intermolecular

    triple hydrogen bonding interactions between the melamine

    and cyanuric acid functionalities.19 In nonpolar solvents

    such as methylcyclohexane, the perylene bisimide unit

    in   14   stacked strongly to form filamentous precipitates.

    FE-SEM revealed that the filaments were composed of 

    intertwined thinner fibrils (Figure 10a and b), while a

    magnified image showed that the fibrils had ribbon-likemorphology with widths of about 100nm (Figure 10c),

    which has been attributed to the ordered stacking of the

    perylene bisimide unit. In contrast, SEM images revealed

    that a quadruple hydrogen-bonding-driven supramolecular

    polymer with no aromatic unit in the backbone forms

    cotton-like structures.20

    5 VESICLES

    The formation of vesicles in solution can be characterized

    by DLS. To get more insight into their structures and assem-

    bling mechanisms, their surface morphology is usually also

    studied by using a combination of microscopic methods.

    Zhao et al. have reported that compound  16a  and  b  formed

    supramolecular polymers that were stabilized by hydrazide-

    based quadruple hydrogen bonding motifs.21 In decalin, the

    supramolecular polymers further self-assembled into vesic-

    ular structures. SEM images clearly showed the formation

    of spherical entities of average diameters of about 0.6 and1.0µm, as shown in Figure 11(a) and (b), respectively. The

    vesicles may exhibit defects, which may be regarded as an

    evidence for their hollowness.22 For   16a   and   b, the fluo-

    rescence and TEM micrographs (Figure 11c and d) further

    supported their hollow feature, because obvious luminance

    differences and membranes were observed.

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©  2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    8/13

    8   Techniques

    X

    O

    XO

    N

    OC12H25

    C12H25O

    XO

    OH

    CH3NH CH3

    HOOHOH

    CH3

    O

    N

    C12H25O

    OC12H25

    O

    HN

    OC12H25

    C12H25O

    O

    NH

    12

    13

    X

    X =

    CH3

    OH

    10 kV × 500 500 µm 000045 10 kV × 500.000 0.5 µm 000063

    (a) (b)

    Figure 9   SEM images of the cast-film of the mixture of  12  and13 (1 : 1) on a glass plate. (Reproduced from Ref. 18.  ©  AmericanChemical Society, 2005.)

    Most observed vesicles on the surface are flattened

    because the evaporation of the encapsulated solvents causes

    the ball to collapse. When vesicles are very stable, they may

    retain their three-dimensional shape on the surface, which

    can be observed by SEM.22,23

    Kim   et al. have reportedthat the donor– acceptor complex of   17   and   18   could be

    encapsulated into the cavity of cucurbit[8]uril (CB[8]).23

    The resulting ternary complex  19  formed large vesicles in

    water, as evidenced by SEM and TEM images (Figure 12).

    The SEM image showed large spheres with a diameter

    ranging from 20 nm to 1.2µm and the typical vesicle size

    lied between 400 and 950 nm. The vesicles maintained the

    spherical shape, indicating their robust stability. In contrast,

    the TEM image did not exhibit a comparably clear three-

    dimensional contrast.

    6 NANOTUBES, -PORES, -RIBBONS,-RODS, AND -SPHERES

    The supramolecular self-assembly of organic nanotubes and

    related structures has received increasing interest since the

    discovery of carbon nanotubes in 1991. SEM is a power-

    ful technique for direct observation of their formation. For

    N

    N

    O

    O

    O

    O

    H

    N

    C12H25

    C12H25

    C12H25

    N NH

    NH

    O

    O

    O

    N N

    N NHR′

    NR′′2

    NH

    NN

    NR′HN

    NR′′2

    14: R′ =

      R′′ = C8H17

    Bu

    Et

    15

    N

    N

    O

    O

    O

    O

    N

    NN

    NR′N

    NR′′2

    HH

    NN

    N O

    O

    O

    H

    HN

    N N

    N NR′

    NR′′2

    H H

    N N

    NO

    O

    O

    H

    H

    Supramolecular polymer

    (a) (b)

    (c)

    10 µm 1 µm

    1 µm

    Figure 10   FE-SEM of the filamentous precipitates of thesupramolecular polymer of   14   and   15   formed from methylcy-clohexane (0.3 mM). The length of the bar across the ribbon in(c) is 100 nm. (Reproduced from Ref. 19.  ©  American ChemicalSociety, 2007.)

    example, Bo  et al. have reported that compound   20   self-

    assembled into nanotubes and layered sheets, which were

    driven by   π –π   stacking and hydrogen bonding between

    the amide units.24 The materials were prepared by heat-

    ing its suspension in THF to reflux until all the solidswere completely dissolved and then allowing it to cool

    gradually to room temperature. SEM study of its air-dried

    suspension showed the formation of fibril assemblies with

    a high aspect ratio (Figure 13a). Their open-ended feature

    clearly revealed the tubular structure (Figure 13b). TEM

    images were also utilized to support the presence of the

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©   2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    9/13

    Scanning electron microscopy   9

    O

    O

    NH

    NH

    N

    N

    O

    O

    RO

    RO

    OO

    HN

    O

    HNO

    N

    N

    H

    H

    O

    O

    OR

    OR

    16a: R = CH2CH(CH3)2

    16b: R = n -C8H17

    H

    H

    5 kV × 7,000 2 µm 02/ MAR/09 10 kV × 16,000 1 µm

    10 µm 50 nm2.6 nm

    (a)

    (c) (d)

    (b)

    Figure 11   SEM images of the vesicles of (a) 16a  (1.0mM) and(b)  16b   (1.0 mM); (c) fluorescence micrograph of the vesicle of 16b (20 mM); and (d) TEM image of the vesicle of  16b   (0.4 mM)in decalin. (Reproduced from Ref. 21.  ©   Elsevier, 2010.)

     N NC16H33 C16H33

    C16H33 C16H33

    + +

    HO

    OH

    +

    NN

    N N

    O

    O

    H H

    CH2

    CB[8] =

    =

    N N+ +

    HOOH

    17   18

    19

    (Ternary complex)

    (a) (b)

    500 nm1 µm

    2 µm

    Figure 12   (a) SEM and (b) TEM images of ternary complex19. Samples (6.9 × 10−4 M) were negatively stained with uranylacetate (2 wt% in water) for observation by TEM.23

    nanotubes (Figure 13c and d). Magnified SEM images fur-

    ther revealed the rolled-up style of the nanotubes with a

    NH

    OC11H23

    HN

    O

    C11H23   20

    10 µm 100 nm

    200 nm500 nm

    100 nm 100 nm

    (a)

    (c)

    (e)

    (b)

    (d)

    (f)

    Figure 13   Morphology of   20. (a) SEM image of a sampleprepared by dropping its THF suspension (0.1 mg ml−1)   onto asilicon substrate followed by air drying and coating with Pt;(b) high-magnification SEM image with an open-ended crosssection; (c) TEM image; (d) high-magnification TEM image; (eand f) high-magnification SEM images: internal and externalscrew ends of the self-assembled tubes.24

    scrolling structure (Figure 13e), which was further con-

    firmed by the terminal types of tubes (Figure 13f). SEM

    images were also obtained for samples at different concen-

    trations, which revealed that nanotubes were formed when

    the concentration was low and transformed to layered sheets

    at high concentrations. On the basis of the observations, the

    authors proposed a hierarchical self-assembling mechanism;

    that is, the one-layer nanosheets were first formed, which

    might further roll up to generate the nanotubes or stack to

    give layered sheets.

    Jiang   et al. have utilized SEM to study the morphol-ogy of the aggregates of porphyrin  21.25 The samples were

    prepared by drop-casting a 6 mg ml−1 chloroform solution

    onto the surface of SiO2   substrate or quartz. It is wor-

    thy to note that the obtained films were annealed in a

    chloroform-vapor-saturated desiccator. Only after anneal-

    ing, the molecules self-assemble into long microtubes. The

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©  2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    10/13

    10   Techniques

    21

    N

    NH N

    HN

    C10H21

    C10H21

    C10H21

    C10H21

    a

    b

    c

    a

    cb

    50 nm

    50 nm50 nm

    Figure 14   SEM images of microtube dendrites of   21   fromchloroform-vapor annealing. (Reproduced from Ref. 25.  ©  Amer-ican Chemical Society, 2010.)

    outer diameter of the microtubes ranges from 0.12 to

    6.9µm, and the wall thickness is about 50 nm (Figure 14).

    The microtubes also displayed uniform orientation, sug-

    gesting greater intermolecular order. Most importantly,

    the images showed that the submicrometer-sized tubes

    were often linked together to form Y-junctions, which in

    turn constructed more complicated junctions, that is, the

    dendrites with submicrometer-sized tube branches. These

    submicrometer-sized tube dendrites might further aggre-

    gate to form larger dendrites, the submicrometer-sized tubes

    of which spread out from its root and acted as a branch.

    The assembling mechanism was thus proposed based on

    the SEM observations. Since the microtube dendrites had a

    curved 2D wall with thickness of 50 nm, they were consid-

    ered to be formed by the rolling up of the higher ordered

    multilayered crystalline film through porphyrin stacking,

    which was stabilized by the interdigitation of the alkyl

    chains.

    There are many different forms of nanomaterials, which

    can be assembled from single or two or more differ-ent components. Depending on the self-assembling condi-

    tions, one building block system may generate different

    types of assembled entities. Jiang  et al. reported that com-

    pound   22a   and   b  aggregated into nanoribbons, -spheres,

    or -rods from different solvents,26 as evidenced by SEM

    (Figure 15). The image of  22a  showed that its chloroform

    N

    N N

    N

    CH3COS(CH2)5O O(CH2)5SCOCH3M

    22a: M = 2H

    22b: M = Zn

    (A)

    (C)

    (E) (F)

    (D)

    (B)

    1 µm

    1 µm 100 nm

    10 nm18 nm

    100 nm 100 nm

    1 µm

    e

    Figure 15   SEM images of nanostructures formed by   22a   andb. (A) Air bubbles by  22a  in CHCl3. (B) 3D networks by  22b  inCHCl3. (C) Hollow spheres by  22a  in CH3OH. (D) Rods by  22b

    in CH3OH. (E) Ribbons by   22a  in  n-hexane. (e) Zoom-in imageof the rectangle part in E. (F) Hollow spheres by  22b  in  n-hexane.(f) Zoom-in image of the rectangle part in (F). (Reproduced fromRef. 26.  ©   American Chemical Society, 2008.)

    solution formed a two-dimensionally ordered array of 

    air bubbles with highly monodispersed pores of about

    500 nm in diameter (Figure 15A), reflecting the weakness

    of its intermolecular interaction. In contrast, the chloro-

    form solution of zinc porphyrin   22b   led to the forma-

    tion of three-dimensional network structures (Figure 15B),

    which supported the significant intermolecular interaction

    owing to the Zn–O(=C) coordination bond. When inject-

    ing a small volume of their chloroform solutions intoCH3OH, metal-free porphyrin  22a  self-assembled into hol-

    low spheres (Figure 15C), which was confirmed by the

    two broken spheres, while the self-assembly of   22b   led

    to the formation of rods (Figure 15D) also as a result of 

    the above Zn–O coordination. When   n-hexane was used

    as the medium, SEM revealed a reverse morphology. For

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©   2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    11/13

    Scanning electron microscopy   11

    N

    N N

    N

    N

    N

    N

    N Zn

    23

    N

    CH3

    R

    R

    24a: R = H

    24b: R = t -Bu

    300 nm

    500 nm

    300 nm

    300 nm

    500 nm

    1 µm

    2 µm

    5 µm

    2.5 µm

    3 µm

    60

    40

    20

    0

       N  u  m

       b  e  r

    7.82 ± 2.07 µm

    (a) (b)

    (c) (d)

    (e) (f)

    2 4 6 8 10 12 14

    Length (µm)

    Figure 16   SEM images of (a)  23   tube, (b) C60 –23   rod,(c) C70 –23   rod, (d)  23 –24a   rod, and (e) 23 –24b   rod. (f) Thelength distribution of   23–24b   rod. (Reproduced from Ref. 27.©   American Chemical Society, 2009.)

    22a, ribbons of uniform size and orientation (Figure 15E),

    with an average width of about 19 nm estimated from the

    zoom-in image (Figure 15e), were assembled. In contrast,the self-assembly of  22b  gave hollow spheres (Figure 15F),

    which was also evidenced by the broken spheres on the

    zoom-in image (Figure 15f).

    Sandanayaka  et al. have used SEM to study the surfactant

    (cetyltrimethylammonium bromide)-assisted self-assembly

    of   23   in the presence and absence of C 60, C70, and   24a

    and   b   in the mixtures of DMF and acetonitrile.27 It was

    revealed that all the samples formed nanotubes or nanorods,

    but their sizes and structural characteristics were very

    different. The zinc porphyrin itself produced tubes with

    a large hollow hole (Figure 16a). In the presence of the

    fullerenes, the hollow holes were completely closed to give

    rods (Figure 16b–e), indicating that the fullerenes were

    effectively encapsulated within the porphyrin tube. SEM

    also helped to produce a detailed distribution of length and

    diameter for the rods, as shown in Figure 15(F) for the rods

    of 23–24b. The average sizes of the rods of the complexes

    were estimated to be 15, 20, 60, and 80 nm, respectively.

    The larger sizes of the rods of  24a  and  b  suggested that an

    encapsulation process of the fullerenes by the tube of   23

    occurred after injecting. Similar to that of 23   (Figure 16a),

    the cross-sectional shape of the   23–C60   and   23–C70   rods

    (Figure 16b and c) were hexagonal. In contrast, the  23 –24a

    and   23– 24b   rods (Figure 16d and e) adopted a distorted

    polygonal shape. The difference has been ascribed to

    the sizes of the nanoparticles. With increasing sizes of fullerenes, the relative size of the flake assembly of   23

    decreased. As a result, the macroscopic organization of the

    assemblies of   23   in the   23–24a   and   23–24b   rods might

    hamper the formation of hexagonal structures, leading to

    the distorted structures.

    7 SUPRAMOLECULAR CHIRALITY

    In solution and gel state, supramolecular chirality is mainly

    investigated by using the circular dichroism spectroscopy. If 

    the samples form ordered microstructures, the helical chiral-

    ity may be expressed and observed on the surface by SEM

    or other microscopic methods. Shinkai  et al. have reported

    that the mixture of compound   25a   and   b   gelated acetic

    acid.28 When tetraethoxylsilane polymerization was carried

    out in this gel (R  =   [25b / 25a  +  25b]   = 5– 15 mol%) and

    the resulting polymer was calcinated, SEM showed that the

    obtained silica not only retained the fibrous structure of 

    the gel but also possessed a right-handed helical structure

    (Figure 17), characteristic of the supramolecular assemblies

    of a chiral molecule. Since the inner diameter (about 10 nm)

    of these helical fibers, estimated by TEM, was comparable

    with that of the gel fibers, it was proposed that the chirality

    in the organogel fibers was transcribed into these inorganic

    silica fibers.

    The above gels of compound   25a   and   b   do not formhelical fibers that can be observed by SEM. However,

    several chiral gelators do selectively self-assemble into

    helicoidal fibers. For example, Escuder  et al. have reported

    that (S,S )-26  gelated several organic solvents.29 SEM of 

    the gel formed in benzene showed the presence of isolated

    right-handed twisted ribbons of several micrometers of 

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©  2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    12/13

    12   Techniques

    NN

    O

    ORO

    25a: R = (CH2)2CH3

    25b: R = (CH2)4N(CH3)3+Br−

    500 nm

    250 nm

    (a)

    (b)

    Figure 17   SEM images of silica structures prepared using gel

    of   25a   and   25b   as a template: (a) SEM for   R  = 25 mol% and(b) SEM for  R  = 10 mol%. (Reproduced from Ref. 28.  ©  RoyalSociety of Chemistry, 2001.)

    length and pitch, together with cylindrical objects and

    longer fibers (Figure 18). SEM also revealed that, when the

    gel was formed by fast cooling of the hot solution, more

    helicoidal fibers were observed, but their sizes decreased.

    In another study, Escuder   et al. have prepared (R,R)-

    26.30 SEM images showed that it produced helicoidal fibers

    of opposite handedness in its benzene gel (Figure 19).

    The SEM images of the precipitates of their racemic

    and nonracemic mixtures were also recorded, which did

    not show the presence of helicoidal aggregates. Since theself-assembly of the fibers is affected by many factors,

    it is nearly impossible to produce and observe fibers of 

    opposite handedness that have perfectly identical shape and

    size. Thus, the above SEM results just illustrate that the

    molecular chirality can be transferred and amplified to the

    whole supramolecular system.

    NH HN

    HNO

    HN O

    (S ,S )-26

    NH HN

    HNO

    HN O

    (R ,R )-26

    Figure 18   SEM picture of (S,S )-26, showing the right-handedhelices formed in benzene. (Reproduced from Ref. 29.  ©  RoyalSociety of Chemistry, 2002.)

    3 µm 2 µm

    (a) (b)

    Figure 19   SEM images showing the details of the helicoidalfibers found in the benzene gels of (S,S )-26   (a) and (R,R)-26(b).30

    8 CONCLUSION

    SEM is a versatile technique for supramolecular science

    to elucidate the microscopic structures of self-assembled

    systems owing to its high lateral resolution and great depth

    of focus. SEM is not in competition with other microscopic

    techniques as it allows different imaging modes. In many

    cases, it is used together with other microscopic techniques.

    Since it covers the nano- to microscales, it is particularly

    useful for observing the morphologies of the self-assembledsystems, but cannot be used to characterize the single

    “pure” supermolecules such as rotaxane, catenane, knot,

    and dendrimer.

    Since the imaging is performed under high vacuum

    for dried samples, conventional SEM cannot avoid struc-

    tural distortion of the studied samples. Thus, the observed

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©   2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044

  • 8/15/2019 smc044

    13/13

    Scanning electron microscopy   13

    morphologies are actually those of the collapsed samples

    without solvents in them. To overcome this limitation, a

    new scanning transmission electron microscopy (STEM)

    imaging technique has been developed, which allows trans-

    mission observations of wet samples in an ESEM.31 How-

    ever, its application for supramolecular science is not

    reported yet.

    REFERENCES

    1. L. Reimer and H. Kohl,   Scanning Electron Microscopy:Physics of Image Formation and Microanalysis, Springer,Berlin, 2008, p. 570.

    2. P. Echlin,   Handbook of Sample Preparation for Scanning Electron Microscopy and X-Ray Microanalysis, Springer,Berlin, 2009, p. 330.

    3. A. Bogner, P.-H. Jouneau, G. Thollet, et al.,  Micron, 2007,38, 390.

    4. M. Knoll, Zeit. Tech. Phys.  1935,  16 , 467.5. V. Seydewitz, Scanning electron microscopy (SEM), in Elec-

    tron Microscopy of Polymers, ed. G. H. Michler, Springer,Berlin, 2008, pp. 87–120.

    6. G. D. Danilatos, Microsc. Res. Tech., 1993,  25 , 529.

    7. R. P. Apkarian, Cryo-temperature stages in nanostructuralresearch, in   Scanning Microscopy for Nanotechnology,eds. W. Zhou and Z. L. Wang, Springer, Berlin, 2007,pp. 467–490.

    8. P. Terech, J. J. Allegraud, and C. M. Garner,   Langmuir ,1998,  14 , 3991.

    9. E. Krieg, E. Shirman, H. Weissman,   et al.,   J. Am. Chem.Soc., 2009,  131 , 14365.

    10. C. Wang, D. Zhang, and D. Zhu,  J. Am. Chem. Soc., 2005,

    127, 16372.11. P. Chen, R. Lu, P. Xue, et al.,  Langmuir , 2009,  25 , 8395.

    12. W. Cai, G.-T. Wang, P. Du, et al., J. Am. Chem. Soc., 2008,130, 13450.

    13. J. G. Hardy, A. R. Hirst, D. K. Smith,   et al.,   Chem. Com-mun., 2005, 385.

    14. M. Enomoto, A. Kishimura, and T. Aida, J. Am. Chem. Soc.,2001,  123 , 5608.

    15. Y. Wang, L. Tang, and J. Yu,  Cryst. Growth Des., 2008,   8,884.

    16. S. Yagai, T. Kinoshita, M. Higashi, et al., J. Am. Chem. Soc.,2007,  129 , 13277.

    17. B. Jancy and S. K. Asha, Chem. Mater., 2008,  20 , 169.

    18. T. Haino, Y. Matsumoto, and Y. Fukazawa,   J. Am. Chem.Soc., 2005,  127 , 8936.

    19. S. Yagai, Y. Monma, N. Kawauchi, et al.,   Org. Lett., 2007,9, 1137.

    20. H. M. Keizer, R. P. Sijbesma, J. F. G. A. Jansen,   et al., Macromolecules, 2003,  36 , 5602.

    21. P. Du, G.-T. Wang, X. Zhao, et al.,  Tetrahedron Lett., 2010,51, 188.

    22. L. Wang, Z.-Y. Xiao, J.-L. Hou,   et al.,   Tetrahedron, 2009,65, 10544.

    23. Y. J. Jeon, P. K. Bharadwaj, S. W. Choi,   et al.,   Angew.Chem. Int. Ed., 2002,  41 , 4474.

    24. Y. Chen, B. Zhu, F. Zhang,   et al.,   Angew. Chem. Int. Ed.,2008,  47 , 6015.

    25. P. Ma, Y. Chen, Y. Bian, and J. Jiang, Langmuir , 2010,  26 ,3678.

    26. Y. Gao, X. Zhang, C. Ma, et al.,   J. Am. Chem. Soc., 2008,130, 17044.

    27. A. S. D. Sandanayaka, T. Murakami, and T. Hasobe, J. Phys. Chem. C , 2009,  113 , 18369.

    28. Y. Ono, K. Nakashima, M. Sano,   et al.,   J. Mater. Chem.,2001,  11 , 2412.

    29. J. Becerril, M. I. Burguete, B. Escuder,  et al.,  Chem. Com-mun., 2002, 738.

    30. J. Becerril, B. Escuder, J. F. Miravet,   et al.,   Eur. J. Org.Chem., 2005, 481.

    31. A. Bogner, G. Tholleta, D. Bassetb, et al., Ultramicroscopy,2005,  104 , 290.

    Supramolecular Chemistry: From Molecules to Nanomaterials, Online  ©  2012 John Wiley & Sons, Ltd.

    This article is  ©  2012 John Wiley & Sons, Ltd.

    This article was published in the  Supramolecular Chemistry: From Molecules to Nanomaterials  in 2012 by John Wiley & Sons, Ltd.

    DOI: 10.1002/9780470661345.smc044