Next Article in Journal
Crystal Structure and Thermal Behavior of SbC2O4OH and SbC2O4OD
Next Article in Special Issue
A Chiral Bis(salicylaldiminato)zinc(II) Complex with Second-Order Nonlinear Optical and Luminescent Properties in Solution
Previous Article in Journal
Metal–Dithiolene Bonding Contributions to Pyranopterin Molybdenum Enzyme Reactivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polymer Coated Semiconducting Nanoparticles for Hybrid Materials

Department of Chemistry, University of Mainz, Duesbergweg 10-14, D-55128 Mainz, Germany
Submission received: 22 January 2020 / Revised: 25 February 2020 / Accepted: 5 March 2020 / Published: 11 March 2020
(This article belongs to the Special Issue Hybrid Inorganic-Organic Luminescent Materials)

Abstract

:
This paper reviews synthetic concepts for the functionalization of various inorganic nanoparticles with a shell consisting of organic polymers and possible applications of the resulting hybrid materials. A polymer coating can make inorganic nanoparticles soluble in many solvents as individual particles and not only do low molar mass solvents become suitable, but also polymers as a solid matrix. In the case of shape anisotropic particles (e.g., rods) a spontaneous self-organization (parallel orientation) of the nanoparticles can be achieved, because of the formation of lyotropic liquid crystalline phases. They offer the possibility to orient the shape of anisotropic nanoparticles macroscopically in external electric fields. At least, such hybrid materials allow semiconducting inorganic nanoparticles to be dispersed in functional polymer matrices, like films of semiconducting polymers. Thereby, the inorganic nanoparticles can be electrically connected and addressed by the polymer matrix. This allows LEDs to be prepared with highly fluorescent inorganic nanoparticles (quantum dots) as chromophores. Recent works have aimed to further improve these fascinating light emitting materials.

Graphical Abstract

1. Introduction

The use of nano-objects in the real macroscopic world requires—besides their preparation—also their macroscopic processing, organisation, and orientation. This is because any device in the “real world” is in the micro- to milli-meter range and thus contains many nanoparticles. This requires the homogeneous dispersion of the nano-objects in solvents (aqueous or non-aqueous water based) or organic polymeric materials. However, this is often hindered by massive segregation of the nano-objects from the surrounding matrix [1]. Thus, concepts to improve the compatibility of nano-objects with the surrounding matrix become important (see Section 2). In addition, self-organisation of the nanometre sized objects is attractive in dispersing them in a controlled way and to orient anisotropic nanoparticles in a uniform way. This can be realised, if the nanoparticles become individually dispersed in a mobile matrix allowing their self-organization (as compared to agglomerates, in which the multitude of interactions hinder a reorientation). It can be achieved in colloidal crystals [2,3,4] and lyotropic (solvent induced) liquid crystals (LC) [5,6] (see Section 3). At least the topic of a “functional coating” becomes important and will be discussed in Section 4 (QD-LEDs), where the focus is put on semiconducting properties. In combination with semiconducting polymers as a matrix it is then possible to address the nanoparticles electrically and to prepare quantum dot LEDs (that are polymeric LEDs with quantum dots as chromophores). So, altogether concepts to achieve a solubilization of nanoparticles (Section 2), their organization (Section 3), and the successive electrical addressing of fluorescent nanoparticles (Section 4) are the topic of this review.

2. Concepts for the Functionalization of Inorganic Nanoparticles to Improve Their Solubility and Processability

In this context it is advisable to start by considering the general concepts for the stabilization of colloids in a matrix [7,8,9,10]. This relies on either electrical charges or a steric stabilization to overcome the strong adhesion forces among the nanoparticles [11,12,13] (see Figure 1). These concepts are, in addition, already used in “nearly any” synthesis of nanoparticles, which is performed in the presence of detergents, to prevent aggregation and to stabilize the nanoparticles formed. In this context an electrostatic stabilization is most often used, which can result from an excess of oxides or cations in the nanoparticle. However, for applications in combination with electric fields (discussed later) it is advisable to use steric stabilization and not electrostatic stabilization to avoid water as solvent. In addition, electrostatic stabilization is less effective for a dispersion in polymeric materials. This makes concepts for a steric stabilization, like “hairy rods” most attractive. Even though this concept was developed for stiff main chain liquid crystalline polymers, it is also applicable for inorganic nanoobjects. [14] The concept implies, that a stiff insoluble core is solubilised by linking long chains (“hair”) onto its surface, which are soluble in the solvent. As “hair” are often used alkyl chains [15,16,17], which can stabilize sphere- or rod-like shaped nanoparticles. They are also in use for delaminated clay fragments [18,19], which possess a plate-like shape.
Polymeric surfactants [7,9,18,19,20,22,23] offer here advantages for the solubilisation of inorganic nano-objects because polymers themselves are objects of nanometre dimensions (for comparison: the alkyl chain in a detergent may possess 18 C–C bonds, while a polymer chain of a degree of 100 has 200 C–C bonds, which bring it, even in the coiled conformation, to the size of many nms). Therefore, polymers for steric stabilization match the size of inorganic nano-objects much better than simple alkyl chains. Thus, polymer coated surfaces are stabilised sterically up to a distance of nanometres and not only for angstroms as with alkyl chains. In addition, polymer chains have, in the coiled conformation, a lot of free volume inside, which can be filled by solvent. This large interaction volume between solvent and functionalised surface leads to good solubility in solvents for the polymer.
As result it is possible to disperse inorganic nanoparticles as individual objects in various solvents, which are good solvents for the polymers used for coating [20,24], which applies also to the use of compatible polymers. It is thus possible to disperse them homogeneously in polymer films as “solid” solvents. The preparation of highly concentrated solutions also becomes possible. This can finally open the door for the incorporation of fluorescent nanoparticles in semiconducting polymers and for their electrical addressing [24,25].
To coat inorganic nanoparticles with a dense and stable layer of polymer chains two strategies exist, which are either grafting to or grafting from [26,27,28,29,30,31,32,33,34,35] (see Figure 2). From a polymer perspective the grafting to approach offers here the advantage that the polymer chains used for grafting can be characterized very well. By the use of multiple anchor groups it becomes possible to avoid adsorption–desorption equilibria and to achieve a stable fixation.
Recently a block copolymer synthesis via RAFT polymerisation in combination with a reactive ester intermediate to introduce the anchor groups attracted much interest [20,21] (see Figure 3). It can be used to introduce different anchor structures to the same block-copolymer, which are suitable for grafting to oxidic or chalkogenidic nanoparticles [36,37]. This is important because oxidic nanoparticles attract much attention because of their magnetic or semiconducting properties, whereas nanoparticles based on selenides (often with a sulfidic shell) are prominent as fluorescent nanoparticles. In addition the new polymers can be designed to have multiple anchor units for surface attachment. Through the multi-dentate fixation of the polymer onto the surface adsorption–desorption equilibria reactions are avoided. In this way not only sphere- or rod-like nanoparticles, but also more complex structures like tetrapods [38] have been successfully coated and dispersed.
Such inorganic nanoparticles coated with polymer chains can be dispersed as indivual nano-objects [7,9,39,40]. This can be seen e.g., from AFM- and TEM-measurements of this films (see Figure 4A). AFM makes it thereby also possible to differentiate between the inorganic core (hard) and the softer polymer corona and to prove thus directly the core–shell structure of the hybrid materials. The direct proof that there are individually dispersed nano-objects comes mostly from light scattering experiments [39,41] (see Figure 4B). Here both dynamic light scattering, which measures the hydrodynamic radius, and static light scattering, which measures the radius of gyration, give similar results, which correspond to the size of the individual object. Some differences between both measuring technics are, however, to be expected, because the hydrodynamic radius (dynamic light scattering) includes the size of the swollen polymer brushes, whereas the radius of gyration is strongly dominated by the inorganic core, which has a higher refractive index increment (see [39] for a detailed discussion).
During the last decade, remarkable progress has been made in colloidal synthesis [42,43,44,45,46] and the polymer coating [47,48,49,50,51,52] of nanoparticles with well controlled shapes, which range from simple dots [53,54,55,56,57] to rods [58,59,60,61] and to more complicated structures such as tetrapods [62,63,64,65], hyperbranched structures [66,67,68], and wires [69,70,71] among others. For most of these geometrical structures the solubilization enhancing effect of a polymer coating has already been shown. A large variety of semiconductor materials, e.g., metal sulfides [11,12,13], metal selenides [72,73,74], metal tellurides [75,76], metal oxides [77,78,79] and lead halide perowskites [80] have been investigated in relation to this topic. They can be coated using the anchor groups presented in Figure 3.
Depending on the solubility of the polymer chains it is also possible to prepare nano-particle solutions, which change their solubility reversible with temperature [39] (see Figure 5a). Alternatively, it becomes possible to include photocleavable links to the polymer chain and to split-off the solubilizing polymer chains on irradiation [81] (see Figure 5b). The solubility can also be changed after preparation of the solution by external stimuli. This might be especially interesting for the dispersion of nanoparticles in polymer matrices, because their viscosity is high and thus solution and dissolution is slow. It is thus possible to stabilize intermediate states of dispersion like finely percolated structures [24], which are interesting for photovoltaics (see below for a more detailed discussion)
Semiconducting polymers (mostly conjugated polymers) and fluorescent inorganic nanoparticles have been investigated separately for some time. The combination of both materials is, however, highly interesting as it can lead to nanocomposites whose properties are much superior to those of their constituents [82]. These hybrid materials are being considered in several applications, mostly in light emitting diodes and photovoltaic cells. Latest reviews about this topic can be found in [83,84,85,86,87]. Semiconducting hybrid systems offer the potential to combine the benefits of inorganic nanocrystals (NCs) such as tunable optical properties, optical stability, and high electron mobility [82,86] with the advantages of semiconducting polymers like light weight, flexibility, ink jet printing, roll-to-roll production, solvent processability over a large area and with low costs [83,88,89,90,91,92]. Because it is possible to tune the optical band gap of NCs very accurate from the ultraviolet to the near-infrared, they can be used to prepare very bright quantum dot light emitting diodes (QLEDs, see Section 4) or solar cells consisting of p-type donor polymers and n-type acceptor inorganic nanoparticles which absorb a significant range of the solar spectrum.
Controlling the morphology is a crucial aspect for optoelectronic applications. For QLEDs a homogenously dispersed structure is most desirable. Therefore, intimate contact between NCs and polymers (no demixing) is essential. Preparation techniques such as kinetic entrapment [93] and in situ polymerization [94] are known to generally stabilize highly dispersed states in nanocomposites.
Solar cells need a percolated structure to facilitate the macroscopic transport of holes and electrons to the appropriate electrode, even though they benefit from improved dispersion (short exciton diffusion lengths). To prepare a percolated structure a controlled demixing is required, which must be stopped (to prevent further coarsening) before the demixed structures become larger than several hundred nms (see [23] for discussion). In this context the concepts discussed in the context of Figure 5 [39,81] are suitable to induce a demixing.

3. Controlling Orientation, Dispersion, and Percolation of Inorganic Nanoparticles in a Polymer Matrix

As discussed in the first paragraph, polymer chains can be used to cover inorganic nanoparticles, which makes them well soluble as individual objects and allows the preparation of highly concentrated solutions even in a polymer matrix. Through this, it is possible to fill polymeric materials with a high concentration of inorganics [7,39,40,41,95]. Furthermore, even functional inorganic nanoparticles can be combined with a functional semiconducting polymer matrix.
In the case of shape anisotropic inorganic nano-objects their good solubility can also give rise to the formation of liquid crystalline phases. That is because above certain “critical” concentrations [20,21,39] the shape anisotropic nano-objects act as mesogens and self-organize [6,7]. This can be also attractive for opto-electronic applications. Generally, the formation of lyotropic liquid crystals [14,96,97] from rigid-rod objects as result of form-anisotropy is well understood [98,99] and such phases offer the potential to orient anisotropic nanoparticles.
Historically lyotropic LC-phases in water have been observed for various rigid-rod objects like V2O5 ribbons [100,101], tobacco mosaic viruses [102], and TiO2 nanorods [103,104]. Aqueous solutions offer the opportunity to stabilise nano-objects electrostatically by surface charges. Thus lyotropic (solvent induced) liquid crystalline phases from V2O5 solutions in water have been known for a long time [22] under proper pH and salt conditions. However, the charges used to stabilise the system require water as solvent and disable any electronic use. Therefore, ion free mineral liquid crystals from uncharged anisotropic nanoparticles in organic solvents are attractive and can be made following the concept of grafting to polymer chains to the anisotropic object.
Using this concept, smectic and nematic phases could be prepared recently in highly concentrated solutions of TiO2 nanorods [20], which were coated with block copolymers (see Figure 6). Polarizing optical microscopy showed birefringent, mobile textures, while TEM measurements of samples prepared from concentrated solutions showed the parallel orientation of the rod like nanoparticles, which are the shape anisotropic mesogens of the system. This work has been extended to various other semi conducting nano-rods from oxidic and chalkogenidic semiconductors like TiO2, ZnO, SnO2 or CdSe, ZnS, and CdTe [21], which required different anchor blocks (see Figure 3). This concept was also applied successfully to induce a liquid crystalline orientation in dispersions of carbon nanotubes [105,106,107]. A comparable ordering of plate like nanoparticles can also be obtained and is used to make artificial nacre [95].
Generally, the possibility to orient shape anisotropic nanoparticles and especially semiconducting ones is interesting for materials science and especially for photovoltaics [108,109,110,111]. In this context Alivisatos and co-workers [15] could recently show that rod-shaped nanoparticles oriented perpendicular to the electrodes improving the efficiency of solar cells [112]. In their case they had to orient the semiconducting inorganic nanorods however by a complicated growing process, away from the substrate. Liquid-crystalline phases, would allow an orientation of the whole phase. For non-properly coated nanoparticles, it is, however, difficult to obtain the high concentrations needed to obtain them. The situation is, however, different for the systems and the polymer coating discussed here.
Recently it was found that polymer coated TiO2 nanorods could be solubilised and oriented also in an organic hole conducting matrix where they orient into a liquid crystalline phase by self-assembly [113]. Moreover, it is possible to orient this liquid crystalline phase uniformly in an external electric field. [114] Thus it becomes possible to orient the central part of the liquid crystalline phase perpendicular to the electrodes (see Figure 7, lower part). The resulting structures have the correct orientation for an organic photovoltaics setup, in which the oriented TiO2 nanorods transport the electrons to one electrode while the holes migrate through the organic hole conducting material.
Two more aspects should be mentioned in the context of the electrical coupling of inorganic nanoparticles and an organic matrix. First, by coating an inorganic nanoparticle with a semiconducting polymer, a p-n junction is established on the level of the core-shell nano-object (see Figure 8). This allows charge transfer and charge migration on the nano-level to be studied [115,116,117]. So it became possible to determine charge separation and charge transport under irradiation for polymer coated hybrid systems [115,116,117]. Figure 8 shows, how the polymer corona charges positively under irradiation, while the inorganic core from TiO2 accumulates electrons (negative charges).
Second, the polymers grafted to the inorganic nanoparticle can be a precursor for graphitic material, into which they decompose during heating (see Figure 9a). This allows the coating of redoxactive inorganic nanoparticles with graphite or, if the process is done in a composite, the dispersion of many nanoparticles in a graphitic matrix. This can be beneficial to electrically address nanoparticular electrode materials in lithium or sodium ion batteries [118,119,120] (see Figure 9) where it helps to improve the electrode performance.
So, to summarize the result of Section 3: The high solubility in “classical” solvents and polymers (as solid solvents), which can be achieved with the polymer coating of inorganic nanoparticles, allows liquid crystalline phased rom shape anisotropic nanoparticles to be prepared easily. These phases can then be used to orient the nanoparticles macroscopically by an external electric field. In addition, the coating with functional polymers makes an electrical coupling to an external matrix possible, this opens up the possibility for electric addressing of fluorescent inorganic nanoparticles like quantum dots (QDs). This will be discussed in Section 4.

4. QD-LEDs

Generally, in semiconducting nanoparticles the electronic properties (e.g., band gap) depend on their size. This is especially obvious for fluorescent nanoparticles, for which the wavelength of fluorescence varies with size. Such fluorescent nanoparticles are called quantum dots (QDs). They possess often a high fluorescence efficiency. So, they are attractive as chromophores in LEDs. Thus recently, quantum dot based light emitting diodes (QLEDs) became competitive alternatives to organic light emitting diodes (OLEDs) in terms of colour purity, luminescence intensities, and external quantum efficiencies (EQEs) [121,122,123] The narrow emission profile, the high stability of quantum dots (QDs), the high photoluminescence quantum yields (PL QYs), and the easily tunable emission wavelengths make them attractive as quantum dot LEDs (QLED) [124]. By understanding the basic device physics and optimizing the device structure it was possible to improve the device performance and efficiency of QLEDs, so much that they can be compared to organic LEDs [122,125,126,127,128,129,130,131,132,133]. For this study semiconducting polymer hybrids QD, which allow a homogeneous dispersion in the matrix of a semiconducting polymer and an efficient charge transport into the QDs were very helpful [116,134,135,136].
Chemical structures of a combination of semiconducting polymers and QDs are presented in Figure 10.
The strong affinity between the surface of the inorganic part and the anchoring block of the polymers leads to a dense coating of the QDs and this allows a good dispersion (uniform distribution) within the semiconducting polymer matrix (Figure 11a). In addition to that the electro-luminescence observed in such hybrid systems is that of the pure QD (Figure 11b). This proves that all excimers formed from the positive and negative charges in the polymer matrix reach the QD for emission. With these systems it becomes possible to study the influence of the morphology of hybrid emission layers (e.g., QD concentrations or mean inter-dot distances) simply by varying the mixing ratios between the QDs and semiconducting polymers (see Figure 11) [24,135]. Using different semiconducting polymers, the charge carrier transport through the semiconducting polymers can be adapted. This allows us to raise the mean QD-to-QD distance or the thickness of the QD active layers without declining the charge transport properties [137,138].
In this context it becomes also attractive to use shape anisotropic fluorescent nanoparticles (see above) to optimize emission [139].
To optimize such QLEDS concerning the external quantum efficiency (EQE), CdSe/CdxZn1−xS core/shell Type-I heterostructured QDs with a core radius of 2.0 nm and a total radius of 4.5 nm were used. Such a structure confines both hole and electron wave functions within the CdSe core. In addition, this specific type of QD was extensively used in previous QLED and QD-semiconducting polymer hybrid studies [127,138]. For further improvement it is necessary to consider the band gap alignment as presented in Figure 12. Crucial for QLEDS with CdSe QDs are especially their low HOMO and LUMO values. They make it, on one hand, difficult to transport positive charges (holes) into the QD and lead, on the other hand, to an easy overcharging of the QDs with electrons. This general problem can be reduced either by optimizing the setup and the injection layers [140,141], or, in the case of the homogeneous dispersion in bulk, by lowering the HOMO level of the conductive polymers grafted to the QDs [24,137,138]. Figure 12 shows the energetic situation (band gap position of the sublayers in the QLED) together with some polymer structures recently synthesized to optimize the transport of positive and negative charges.
Recently the significance of the morphology of QD emission layer on the device performance was studied in a systematic comparison [25] and it could be shown that the homogeneous dispersion of QDs, as it can be obtained with hybrid systems, is advantageous. For this purpose, QD emission layers of different morphologies consisting of QD-only films, QD/semiconductor polymer blends or QD-semiconducting polymer hybrids, were compared (see Figure 13). The optoelectronic performance of QLEDs demonstrates that the hybridization of QDs with semiconducting polymers significantly enhances the emission efficiency of QD emitters and this can be explained by (i) the improved charge carrier balance concerning the QDs and (ii) the optimized morphology. Pure blends suffer from their inhomogeneous structure due to segregation of QDs and polymer (see also Figure 11). Thus, in parts of the film the charges will never meet a QD, while in others the situation resembles QD-only films. QD-only films are, however, very fragile and cannot be expanded in thickness; they bring the QDs in close proximity to each other, which increases the electronic coupling between the QDs and leads to quenching. Thus, the efficiency of the QD emission layer can be increased by increasing the mean inter-dot distance as happens naturally in the hybrid QD emission layers. This oppresses the quenching between QDs. So, QD-semiconducting polymer hybrid emission layers benefit from both aspects and show a substantially enhanced device efficiency (peak EQE of 5.6%) and brightness (peak luminance of 21,707 cd m−2) [25] compared to the cases of QD-only (2.0% and 16,843 cd m−2) or physically blended QD/polymer emission layers (1.7% and 4207 cd m−2).
Such hybrid structures can also be made from InP QDs (see Figure 14). Such QDs have found a lot of interest in circumventing the use of toxic Cd in CdSe. From the chemical side it is relatively easy to coat such InP QDs with semiconducting polymers, because InP quantum dots themselves are usually coated with ZnSe/S as inorganic shell to stabilize the InP core [127]. Therefore, the chemistry needed to graft polymers to them is identical to the established CdSe QD, which uses the same shell material (Figure 14a). In this case successful grafting, homogeneous dispersion of the QDs, and an improved external quantum efficiency could also be demonstrated. The electroluminescence is again identical to that of QD-only devices (Figure 14b).
Quiet recently some new concepts to improve the performance of QD-LEDs by the use of polymer hybrids were introduced. This concept starts with the idea, that the QDs act as chromophores in the semiconducting matrix, which is needed for charge transport. Usually the need to localize the charges (positive and negative) inside the chromophores for excimer formation requires that the QDs have a lower band gap than the matrix. However, under these conditions they will naturally act as traps, reduce the charge carrier mobility [142], and worsen charge transport and effectivity. Now there is a possibility to circumvent this problem. It requires the blocking of charge transfer, because in this way the creation of traps can be eliminated. The transfer of energy from the matrix (excimer formation within the semiconductor) to the chromophore can then still be done by Foerster transfer [142]. This concept could be realized recently [142] (see Figure 15). It was possible by coating of the CdSe QDs with polystyrene. Polystyrene has such a large band gap so that neither electrons nor holes can easily cross a thin shell of it. So, the CdSe QDs no longer act as traps, which improves the charge transport in the matrix of the semiconducting polymer. Furthermore, it leads to a voltage independent electroluminescence spectrum of the quantum dots.

5. Summary

This paper summarizes recent work on the preparation of inorganic–organic hybrids from a materials perspective. For this purpose, it presents, at first, concepts to improve the solubility (or more broadly, the compatibility) of inorganic nanoparticles in organic solvents or polymers as matrix. It has been shown that with a proper polymer coating it is possible to dissolve (disperse) the nanoparticles as individual objects and to modify this,-later on, with external stimuli. The dispersion as individual objects becomes thereby not only possible in solvents, but also in polymers (as solid solvents) and even in highly functional, semiconducting polymers.
Starting with good solubility (compatibility) two further aspects are discussed. The possibility to prepare highly concentrated solutions of shape anisotropic nano-objects make it easy to obtain lyotroic liquid crystalline phases. They also allow a macroscopic orientation of the shape anisotropic nano-objects.
Alternatively, semiconducting inorganic nanoparticles can be electrically connected to a polymer matrix. This is advantageous for energy storage in batteries and it allows polymer LEDs to be prepared with highly fluorescent inorganic nano-particles (QDs) as chromophore. Recent works have aimed to further improve these fascinating light emitting materials.

Funding

Most of the research of the author -in relation to this topic- was funded by the German Science Fundation, DFG (IRTG 1404)

Acknowledgments

Dominik Fuchs is thanked for helping in the preparation of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vaia, R.A.; Wagner, H.D. Framework for nanocomposites. Mater. Today 2004, 7, 32–37. [Google Scholar] [CrossRef]
  2. Bolhuis, P.G.; Kofke, D.A. Monte carlo study of freezing of polydisperse hard spheres. Phys. Rev. E Stat. Phys. Plasmas Fluids Relat. Interdiscip. Top. 1996, 54, 634–643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Bartlett, P.; Warren, P.B. Reentrant melting in polydispersed hard spheres. Phys. Rev. Lett. 1999, 82, 1979–1982. [Google Scholar] [CrossRef]
  4. Fleischhaker, F.; Zentel, R. Photonic crystals from core-shell colloids with incorporated highly fluorescent quantum dots. Chem. Mater. 2005, 17, 1346–1351. [Google Scholar] [CrossRef]
  5. Stegemeyer, H.; Behret, H. (Eds.) Liquid Crystals; Steinkopff: Heidelberg, Germany, 1994. [Google Scholar] [CrossRef]
  6. Demus, D.; Goodby, J.; Gray, G.W.; Spiess, H.-W.; Vill, V. Handbook of Liquid Crystals; Wiley: Hoboken, NJ, USA, 1998. [Google Scholar] [CrossRef]
  7. Fischer, S.; Salcher, A.; Kornowski, A.; Weller, H.; Förster, S. Completely miscible nanocomposites. Angew. Chem. Int. Ed. 2011, 50, 7811–7814. [Google Scholar] [CrossRef] [PubMed]
  8. Israelachvili, J. Intermolecular and Surface Forces, 3rd ed.; Elsevier: Amsterdam, The Netherlands, 2011. [Google Scholar]
  9. Ehlert, S.; Taheri, S.M.; Pirner, D.; Drechsler, M.; Schmidt, H.-W.; Förster, S. Polymer ligand exchange to control stabilization and compatibilization of nanocrystals. ACS Nano 2014, 8, 6114–6122. [Google Scholar] [CrossRef]
  10. Butt, H.-J.; Kappl, M. Surface and Interfacial Forces 2e; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2018. [Google Scholar] [CrossRef]
  11. Bredol, M.; Matras, K.; Szatkowski, A.; Sanetra, J.; Prodi-Schwab, A. P3HT/ZnS: A New hybrid bulk heterojunction photovoltaic system with very high open circuit voltage. Sol. Energy Mater. Sol. Cells 2009, 93, 662–666. [Google Scholar] [CrossRef]
  12. Guchhait, A.; Rath, A.K.; Pal, A.J. To make polymer: Quantum dot hybrid solar cells NIR-active by increasing diameter of PbSnanoparticles. Sol. Energy Mater. Sol. Cells 2011, 95, 651–656. [Google Scholar] [CrossRef]
  13. Ren, S.; Chang, L.Y.; Lim, S.K.; Zhao, J.; Smith, M.; Zhao, N.; Bulović, V.; Bawendi, M.; Gradečak, S. Inorganic-organic hybrid solar cell: Bridging quantum dots to conjugated polymer nanowires. Nano Lett. 2011, 11, 3998–4002. [Google Scholar] [CrossRef]
  14. Ballauff, M. Stiff-Chain Polymers-Structure, Phase Behaviour, and Properties. Angew. Chem. Int. Ed. 1989, 59, 253–267. [Google Scholar] [CrossRef]
  15. Li, L.S.; Walda, J.; Manna, L.; Alivisatos, A.P. Semiconductor nanorod liquid crystals. Nano Lett. 2002, 2. [Google Scholar] [CrossRef] [Green Version]
  16. Jana, N.R. Nanorod shape separation using surfactant assisted self-assembly. Chem. Commun. 2003, 9, 1950–1951. [Google Scholar] [CrossRef] [PubMed]
  17. Lemaire, B.J.; Davidson, P.; Ferré, J.; Jamet, J.P.; Petermann, D.; Panine, P.; Dozov, I.; Stoenescu, D.; Jolivet, J.P. The complex phase behaviour of suspensions of goethite (α-FeOOH) nanorods in a magnetic field. Faraday Discuss. 2005, 128, 271–283. [Google Scholar] [CrossRef] [PubMed]
  18. Herrera-Alonso, M.; McCarthy, T.J. Chemical surface modification of Poly(p-Xylylene) thin films. Langmuir 2004, 20, 9184–9189. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, Z.X.; Van Duijneveldt, J.S. Isotropic-nematic phase transition of nonaqueous suspensions of natural clay rods. J. Chem. Phys. 2006, 124. [Google Scholar] [CrossRef]
  20. Meuer, S.; Oberle, P.; Theato, P.; Tremel, W.; Zentel, R. Liquid crystalline phases from polymer-functionalized TiO2 nanorods. Adv. Mater. 2007, 19, 2073–2078. [Google Scholar] [CrossRef]
  21. Zorn, M.; Meuer, S.; Tahir, M.N.; Khalavka, Y.; Sönnichsen, C.; Tremel, W.; Zentel, R. Liquid crystalline phases from polymer functionalised semiconducting nanorods. J. Mater. Chem. 2008, 18, 3050–3058. [Google Scholar] [CrossRef]
  22. Van Bruggen, M.P.B.; Van Der Kooij, F.M.; Lekkerkerker, H.N.W. Liquid crystal phase transitions in dispersions of rod-like colloidal particles. J. Phys. Condens. Matter 1996, 8, 9451–9456. [Google Scholar] [CrossRef] [Green Version]
  23. Van Der Beek, D.; Reich, H.; Van Der Schoot, P.; Dijkstra, M.; Schilling, T.; Vink, R.; Schmidt, M.; Van Roij, R.; Lekkerkerker, H. Isotropic-nematic interface and wetting in suspensions of colloidal platelets. Phys. Rev. Lett. 2006, 97, 3–6. [Google Scholar] [CrossRef] [Green Version]
  24. Mathias, F.; Fokina, A.; Landfester, K.; Tremel, W.; Schmid, F.; Char, K.; Zentel, R. Morphology control in biphasic hybrid systems of semiconducting materials. Macromol. Rapid Commun. 2015, 36, 959–983. [Google Scholar] [CrossRef]
  25. Fokina, A.; Lee, Y.; Chang, J.H.; Park, M.; Sung, Y.; Bae, W.K.; Char, K.; Lee, C.; Zentel, R. The role of emission layer morphology on the enhanced performance of light-emitting diodes based on quantum dot-semiconducting polymer hybrids. Adv. Mater. Interfaces 2016, 3, 1–9. [Google Scholar] [CrossRef]
  26. Skaff, H.; Ilker, M.F.; Coughlin, E.B.; Emrick, T. Preparation of cadmium selenide-polyolefin composites from functional phosphine oxides and ruthenium-based metathesis. J. Am. Chem. Soc. 2002, 124, 5729–5733. [Google Scholar] [CrossRef] [PubMed]
  27. Javier, A.; Yun, C.S.; Sorena, J.; Strouse, G.F. Energy transport in CdSe nanocrystals assembled with molecular wires. J. Phys. Chem. B 2003, 107, 435–442. [Google Scholar] [CrossRef]
  28. Sill, K.; Emrick, T. Nitroxide-mediated radical polymerization from CdSe nanoparticles. Chem. Mater. 2004, 16, 1240–1243. [Google Scholar] [CrossRef]
  29. Skaff, H.; Emrick, T. Reversible addition fragmentation chain transfer (RAFT) polymerization from unprotected cadmium selenide nanoparticles. Angew. Chem. Int. Ed. 2004, 43, 5383–5386. [Google Scholar] [CrossRef] [PubMed]
  30. Esteves, A.C.C.; Bombalski, L.; Trindade, T.; Matyjaszewski, K.; Barros-Timmons, A. Polymer grafting from CdS quantum dots via AGET ATRP in miniemulsion. Small 2007, 3, 1230–1236. [Google Scholar] [CrossRef]
  31. Zhang, Q.; Russell, T.P.; Emrick, T. Synthesis and characterization of CdSe nanorods functionalized with regioregular Poly(3-Hexylthiophene). Chem. Mater. 2007, 19, 3712–3716. [Google Scholar] [CrossRef]
  32. Sih, B.C.; Wolf, M.O. CdSe nanorods functionalized with thiol-anchored oligothiophenes. J. Phys. Chem. C 2007, 111, 17184–17192. [Google Scholar] [CrossRef]
  33. Sudeep, P.K.; Early, K.T.; McCarthy, K.D.; Odoi, M.Y.; Barnes, M.D.; Emrick, T. Monodisperse oligo (Phenylene vinylene) ligands on CdSe quantum dots: Synthesis and polarization anisotropy measurements. J. Am. Chem. Soc. 2008, 130, 2384–2385. [Google Scholar] [CrossRef]
  34. Huang, Y.-C.; Hsu, J.-H.; Liao, Y.-C.; Yen, W.-C.; Li, S.-S.; Lin, S.-T.; Chen, C.-W.; Su, W.-F. Employing an amphiphilic interfacial modifier to enhance the performance of a Poly(3-Hexyl Thiophene)/TiO2 hybrid solar cell. J. Mater. Chem. 2011, 21, 4450. [Google Scholar] [CrossRef]
  35. Stalder, R.; Xie, D.; Zhou, R.; Xue, J.; Reynolds, J.R.; Schanze, K.S. Variable-gap conjugated oligomers grafted to CdSe nanocrystals. Chem. Mater. 2012, 24, 3143–3152. [Google Scholar] [CrossRef]
  36. Eberhardt, M.; Théato, P. Raft polymerization of pentafluorophenyl methacrylate: Preparation of reactive linear Diblock copolymers. Macromol. Rapid Commun. 2005, 26, 1488–1493. [Google Scholar] [CrossRef]
  37. Eberhardt, M.; Mruk, R.; Zentel, R.; Théato, P. Synthesis of pentafluorophenyl(Meth)Acrylate polymers: New precursor polymers for the synthesis of multifunctional materials. Eur. Polym. J. 2005, 41, 1569–1575. [Google Scholar] [CrossRef]
  38. Lim, J.; Zur Borg, L.; Dolezel, S.; Schmid, F.; Char, K.; Zentel, R. Strategy for good dispersion of well-defined tetrapods in semiconducting polymer matrices. Macromol. Rapid Commun. 2014, 35, 1685–1691. [Google Scholar] [CrossRef] [PubMed]
  39. Meuer, S.; Fischer, K.; Mey, I.; Janshoff, A.; Schmidt, M.; Zentel, R. Liquid Crystals from Polymer-Functionalized TiO 2 nanorod mesogens. Macromolecules 2008, 41, 7946–7952. [Google Scholar] [CrossRef]
  40. Akcora, P.; Liu, H.; Kumar, S.K.; Moll, J.; Li, Y.; Benicewicz, B.C.; Schadler, L.S.; Acehan, D.; Panagiotopoulos, A.Z.; Pryamitsyn, V.; et al. Anisotropic self-assembly of spherical polymer-grafted nanoparticles. Nat. Mater. 2009, 8, 354–359. [Google Scholar] [CrossRef]
  41. Nikolic, M.S.; Olsson, C.; Saldier, A.; Kornowski, A.; Rank, A.; Schubert, R.; Frömsdorf, A.; Weller, H.; Förster, S. Micelle and vesicle formation of amphiphilic nanopartieles. Angew. Chem. Int. Ed. 2009, 48, 2752–2754. [Google Scholar] [CrossRef]
  42. Manna, L.; Scher, E.C.; Paul Alivisatos, A. Shape control of colloidal semiconductor nanocrystals. J. Clust. Sci. 2002, 13, 521–532. [Google Scholar] [CrossRef]
  43. Peng, X. Mechanisms for the shape-control and shape-evolution of colloidal semiconductor nanocrystals. Adv. Mater. 2003, 15, 459–463. [Google Scholar] [CrossRef]
  44. Pellegrino, T.; Manna, L.; Kudera, S.; Liedl, T.; Koktysh, D.; Rogach, A.L.; Keller, S.; Rädler, J.; Natile, G.; Parak, W.J. Hydrophobic nanocrystals coated with an amphiphilic polymer shell: A general route to water soluble nanocrystals. Nano Lett. 2004, 4, 703–707. [Google Scholar] [CrossRef]
  45. Yin, Y.; Alivisatos, A.P. Colloidal nanocrystal synthesis and the organic–Inorganic interface. Nature 2005, 437, 664–670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Zhao, L.; Wang, J.; Lin, Z. Semiconducting nanocrystals, conjugated polymers, and conjugated polymer/nanocrystal nanohybrids and their usage in solar cells. Front. Chem. China 2010, 5, 33–44. [Google Scholar] [CrossRef]
  47. Li, L.S.; Alivisatos, A.P. Semiconductor nanorod liquid crystals and their assembly on a substrate. Adv. Mater. 2003, 15, 408–411. [Google Scholar] [CrossRef] [Green Version]
  48. Ghezelbash, A.; Koo, B.; Korgel, B.A. Self-assembled stripe patterns of CdS nanorods. Nano Lett. 2006, 6, 1832–1836. [Google Scholar] [CrossRef] [PubMed]
  49. Ryan, K.M.; Mastroianni, A.; Stancil, K.A.; Liu, H.; Alivisatos, A.P. Electric-field-assisted assembly of perpendicularly oriented nanorod superlattices. Nano Lett. 2006, 6, 1479–1482. [Google Scholar] [CrossRef] [Green Version]
  50. Querner, C.; Fischbein, M.D.; Heiney, P.A.; Drndić, M. Millimeter-scale assembly of CdSe nanorods into smectic superstructures by solvent drying kinetics. Adv. Mater. 2008, 20, 2308–2314. [Google Scholar] [CrossRef]
  51. Kang, C.C.; Lai, C.W.; Peng, H.C.; Shyue, J.J.; Chou, P.T. 2D self-bundled CdS nanorods with micrometer dimension in the absence of an external directing process. ACS Nano 2008, 2, 750–756. [Google Scholar] [CrossRef]
  52. Zhang, S.Y.; Regulacio, M.D.; Han, M.Y. Self-assembly of colloidal one-dimensional nanocrystals. Chem. Soc. Rev. 2014, 43, 2301–2323. [Google Scholar] [CrossRef]
  53. Brus, L.E. Electron-electron and electron-hole interactions in small semiconductor crystallites: The size dependence of the lowest excited electronic state. J. Chem. Phys. 1984, 80, 4403–4409. [Google Scholar] [CrossRef] [Green Version]
  54. Alivisatos, A.P. Semiconductor clusters, nanocrystals, and quantum dots. Science (80-.) 1996, 271, 933–937. [Google Scholar] [CrossRef] [Green Version]
  55. Dabbousi, B.O.; Rodriguez-Viejo, J.; Mikulec, F.V.; Heine, J.R.; Mattoussi, H.; Ober, R.; Jensen, K.F.; Bawendi, M.G. (CdSe)ZnS core-shell quantum dots: Synthesis and characterization of a size series of highly luminescent nanocrystallites. J. Phys. Chem. B 1997, 101, 9463–9475. [Google Scholar] [CrossRef]
  56. Schwartz, D.A.; Norberg, N.S.; Nguyen, Q.P.; Parker, J.M.; Gamelin, D.R. Magnetic quantum dots: Synthesis, spectroscopy, and magnetism of Co2+- and Ni2+-doped ZnO nanocrystals. J. Am. Chem. Soc. 2003, 125, 13205–13218. [Google Scholar] [CrossRef] [PubMed]
  57. Kamat, P.V. Quantum dot solar cells. Semiconductor nanocrystals as light harvesters. J. Phys. Chem. C 2008, 112, 18737–18753. [Google Scholar] [CrossRef]
  58. Yang, P.; Lieber, C.M. Nanorod-superconductor composites: A pathway to materials with high critical current densities. Science (80-.) 1996, 273, 1836–1840. [Google Scholar] [CrossRef]
  59. Jun, Y.W.; Lee, S.M.; Kang, N.J.; Cheon, J. Controlled synthesis of multi-armed CdS nanorod architectures using monosurfactant system. J. Am. Chem. Soc. 2001, 123, 5150–5151. [Google Scholar] [CrossRef]
  60. Greene, L.E.; Law, M.; Tan, D.H.; Montano, M.; Goldberger, J.; Somorjai, G.; Yang, P. General route to vertical Zno nanowire arrays using textured ZnO seeds. Nano Lett. 2005, 5, 1231–1236. [Google Scholar] [CrossRef]
  61. Pérez-Juste, J.; Pastoriza-Santos, I.; Liz-Marzán, L.M.; Mulvaney, P. Gold nanorods: Synthesis, characterization and applications. Coord. Chem. Rev. 2005, 249, 1870–1901. [Google Scholar] [CrossRef]
  62. Manna, L.; Scher, E.C.; Alivisatos, A.P. Synthesis of soluble and processable rod-, arrow-, teardrop-, and tetrapod-shaped CdSe nanocrystals. J. Am. Chem. Soc. 2000, 122, 12700–12706. [Google Scholar] [CrossRef]
  63. Manna, L.; Milliron, D.J.; Meisel, A.; Scher, E.C.; Alivisatos, A.P. Controlled growth of tetrapod-branched inorganic nanocrystals. Nat. Mater. 2003, 2, 382–385. [Google Scholar] [CrossRef]
  64. Fiore, A.; Mastria, R.; Lupo, M.G.; Lanzani, G.; Giannini, C.; Carlino, E.; Morello, G.; De Giorgi, M.; Li, Y.; Cingolani, R.; et al. Tetrapod-shaped colloidal nanocrystals of II−VI semiconductors prepared by seeded growth. J. Am. Chem. Soc. 2009, 131, 2274–2282. [Google Scholar] [CrossRef]
  65. Lim, J.; Bae, W.K.; Park, K.U.; Zur Borg, L.; Zentel, R.; Lee, S.; Char, K. Controlled synthesis of CdSe tetrapods with high morphological uniformity by the persistent kinetic growth and the halide-mediated phase transformation. Chem. Mater. 2013, 25, 1443–1449. [Google Scholar] [CrossRef]
  66. Avis, C.; Jang, J. High-performance solution processed oxide TFT with aluminum oxide gate dielectric fabricated by a sol-gel method. J. Mater. Chem. 2011, 21, 10649–10652. [Google Scholar] [CrossRef]
  67. Yang, T.; Gordon, Z.D.; Chan, C.K. Synthesis of hyperbranched perovskite nanostructures. Cryst. Growth Des. 2013, 13, 3901–3907. [Google Scholar] [CrossRef]
  68. Li, H.; Kanaras, A.G.; Manna, L. Colloidal branched semiconductor nanocrystals: State of the art and perspectives. Acc. Chem. Res. 2013, 46, 1387–1396. [Google Scholar] [CrossRef]
  69. Lauhon, L.J.; Gudiksen, M.S.; Wang, D.; Lieber, C.M. Epitaxial core–shell and core–multishell nanowire heterostructures. Nature 2002, 420, 57–61. [Google Scholar] [CrossRef]
  70. Milliron, D.; Hughes, S.M.; Cui, Y.; Manna, L.; Li, J.; Wang, L.W.; Alivisatos, A.P. Colloidal nanocrystal heterostructures with linear and branched topology. Nature 2004, 430, 190–195. [Google Scholar] [CrossRef]
  71. Cho, K.S.; Talapin, D.V.; Gaschler, W.; Murray, C.B. Designing PbSe nanowires and nanorings through oriented attachment of nanoparticles. J. Am. Chem. Soc. 2005, 127, 7140–7147. [Google Scholar] [CrossRef]
  72. Dayal, S.; Kopidakis, N.; Olson, D.C.; Ginley, D.S.; Rumbles, G. Photovoltaic devices with a low band gap polymer and CdSe nanostructures exceeding 3% efficiency. Nano Lett. 2010, 10, 239–242. [Google Scholar] [CrossRef]
  73. Lee, J.S.; Kovalenko, M.V.; Huang, J.; Chung, D.S.; Talapin, D.V. Band-like transport, high electron mobility and high photoconductivity in all-inorganic nanocrystal arrays. Nat. Nanotechnol. 2011, 6, 348–352. [Google Scholar] [CrossRef]
  74. Jeltsch, K.F.; Schädel, M.; Bonekamp, J.B.; Niyamakom, P.; Rauscher, F.; Lademann, H.W.A.; Dumsch, I.; Allard, S.; Scherf, U.; Meerholz, K. Efficiency enhanced hybrid solar cells using a blend of quantum dots and nanorods. Adv. Funct. Mater. 2012, 22, 397–404. [Google Scholar] [CrossRef]
  75. Kang, Y.; Park, N.G.; Kim, D. Hybrid solar cells with vertically aligned CdTe nanorods and a conjugated polymer. Appl. Phys. Lett. 2005, 86, 1–3. [Google Scholar] [CrossRef]
  76. Chen, H.C.; Lai, C.W.; Wu, I.C.; Pan, H.R.; Chen, I.W.P.; Peng, Y.K.; Liu, C.L.; Chen, C.H.; Chou, P.T. Enhanced performance and air stability of 3.2% hybrid solar cells: How the functional polymer and CdTe nanostructure boost the solar cell efficiency. Adv. Mater. 2011, 23, 5451–5455. [Google Scholar] [CrossRef] [PubMed]
  77. Beek, W.J.E.; Wienk, M.M.; Kemerink, M.; Yang, X.; Janssen, R.A.J. Hybrid zinc oxide conjugated polymer bulk heterojunction solar cells. J. Phys. Chem. B 2005, 109, 9505–9516. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, X.; Zhuang, J.; Peng, Q.; Li, Y. A general strategy for nanocrystal synthesis. Nature 2005, 437, 121–124. [Google Scholar] [CrossRef]
  79. Lin, Y.Y.; Chu, T.H.; Li, S.S.; Chuang, C.H.; Chang, C.H.; Su, W.F.; Chang, C.P.; Chu, M.W.; Chen, C.W. Interfacial nanostructuring on the performance of polymer/TiO2 nanorod bulk heterojunction solar cells. J. Am. Chem. Soc. 2009, 131, 3644–3649. [Google Scholar] [CrossRef]
  80. Shen, X.; Zhang, Y.; Kershaw, S.V.; Li, T.; Wang, C.; Zhang, X.; Wang, W.; Li, D.; Wang, Y.; Lu, M.; et al. Zn-alloyed CsPbI3 nanocrystals for highly efficient perovskite light-emitting devices. Nano Lett. 2019, 19, 1552–1559. [Google Scholar] [CrossRef]
  81. Mathias, F.; Tahir, M.N.; Tremel, W.; Zentel, R. Functionalization of TiO2 nanoparticles with semiconducting polymers containing a photocleavable anchor group and separation via irradiation afterward. Macromol. Chem. Phys. 2014, 215, 604–613. [Google Scholar] [CrossRef]
  82. Colvin, V.L.; Schlamp, M.C.; Alivisatos, A.P. Light-emitting diodes made from cadmium selenide nanocrystals and a semiconducting polymer. Nature 1994, 370, 354–357. [Google Scholar] [CrossRef]
  83. Reiss, P.; Couderc, E.; De Girolamo, J.; Pron, A. Conjugated polymers/semiconductor nanocrystals hybrid materials—Preparation, electrical transport properties and applications. Nanoscale 2011, 3, 446–489. [Google Scholar] [CrossRef]
  84. Chandrasekaran, J.; Nithyaprakash, D.; Ajjan, K.B.; Maruthamuthu, S.; Manoharan, D.; Kumar, S. Hybrid solar cell based on blending of organic and inorganic materials—An overview. Renew. Sustain. Energy Rev. 2011, 15, 1228–1238. [Google Scholar] [CrossRef]
  85. Zhao, L.; Lin, Z. Crafting semiconductor organic-inorganic nanocomposites via placing conjugated polymers in intimate contact with nanocrystals for hybrid solar cells. Adv. Mater. 2012, 24, 4353–4368. [Google Scholar] [CrossRef] [PubMed]
  86. Wright, M.; Uddin, A. Organic-inorganic hybrid solar cells: A comparative review. Sol. Energy Mater. Sol. Cells 2012, 107, 87–111. [Google Scholar] [CrossRef]
  87. Liu, R. Hybrid organic/inorganic nanocomposites for photovoltaic cells. Materials 2014, 7, 2747–2771. [Google Scholar] [CrossRef] [PubMed]
  88. Arici, E.; Sariciftci, N.S.; Meissner, D. Hybrid solar cells based on nanoparticles of CuInS2 in organic matrices. Adv. Funct. Mater. 2003, 13, 165–170. [Google Scholar] [CrossRef]
  89. Olson, D.C.; Piris, J.; Collins, R.T.; Shaheen, S.E.; Ginley, D.S. Hybrid photovoltaic devices of polymer and Zno nanofiber composites. Thin Solid Film. 2006, 496, 26–29. [Google Scholar] [CrossRef]
  90. Tekin, E.; Smith, P.J.; Hoeppener, S.; Van Den Berg, A.M.J.; Susha, A.S.; Rogach, A.L.; Feldmann, J.; Schubert, U.S. InkJet printing of luminescent cdte nanocrystal-polymer composites. Adv. Funct. Mater. 2007, 17, 23–28. [Google Scholar] [CrossRef]
  91. Krebs, F.C. Fabrication and processing of polymer solar cells: A review of printing and coating techniques. Sol. Energy Mater. Sol. Cells 2009, 93, 394–412. [Google Scholar] [CrossRef]
  92. Kim, J.H.; Park, J.W. Foldable transparent substrates with embedded electrodes for flexible electronics. ACS Appl. Mater. Interfaces 2015, 7, 18574–18580. [Google Scholar] [CrossRef]
  93. Liff, S.M.; Kumar, N.; McKinley, G.H. High-performance elastomeric nanocomposites via solvent-exchange processing. Nat. Mater. 2007, 6, 76–83. [Google Scholar] [CrossRef]
  94. Lu, Y.; Yang, Y.; Sellinger, A.; Lu, M.; Huang, J.; Fan, H.; Haddad, R.; Lopez, G.; Burns, A.R.; Sasaki, D.Y.; et al. Self-assembly of mesoscopically ordered chromatic polydiacetylene/silica nanocomposites. Nature 2001, 410, 913–917. [Google Scholar] [CrossRef]
  95. Das, P.; Malho, J.-M.; Rahimi, K.; Schacher, F.H.; Wang, B.; Demco, D.E.; Walther, A. Nacre-mimetics with synthetic nanoclays up to ultrahigh aspect ratios. Nat. Commun. 2015, 6, 5967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Han, H.; Bhowmik, P.K. Wholly aromatic liquid-crystalline polyesters. Prog. Polym. Sci. 1997, 22, 1431–1502. [Google Scholar] [CrossRef]
  97. Davidson, P.; Gabriel, J.C.P. Mineral liquid crystals. Curr. Opin. Colloid Interface Sci. 2005, 9, 377–383. [Google Scholar] [CrossRef]
  98. Flory, P.J.; Ronca, G. Theory of systems of rodlike particles—1. Athermal systems. Mol. Cryst. Liq. Cryst. 1979, 54, 289–309. [Google Scholar] [CrossRef]
  99. Vlassopoulos, D. Determination of chain conformation of stiff polymers by depolarized rayleigh scattering in solution. Macromolecules 1996, 29, 8948–8953. [Google Scholar] [CrossRef]
  100. Zocher, H. Über freiwillige Struktur Bildung in Solen. (Eine neue Art anisotrop flüssiger Medien.). Z. für Anorg. Allg. Chem. 1925, 147, 91–110. [Google Scholar] [CrossRef]
  101. Wang, Y.; Takahashi, K.; Lee, K.; Cao, G. Nanostructured vanadium oxide electrodes for enhanced lithium-ion intercalation. Adv. Funct. Mater. 2006, 16, 1133–1144. [Google Scholar] [CrossRef]
  102. Wegner, S.; Börzsönyi, T.; Bien, T.; Rose, G.; Stannarius, R. Alignment and dynamics of elongated cylinders under shear. Soft Matter 2012, 8, 10950–10958. [Google Scholar] [CrossRef]
  103. Dessombz, A.; Chiche, D.; Davidson, P.; Panine, P.; Chanéac, C.; Jolivet, J.P. Design of liquid-crystalline aqueous suspensions of rutile nanorods: Evidence of anisotropic photocatalytic properties. J. Am. Chem. Soc. 2007, 129, 5904–5909. [Google Scholar] [CrossRef]
  104. Michot, L.J.; Bihannic, I.; Maddi, S.; Baravian, C.; Levitz, P.; Davidson, P. Sol/Gel and isotropic/nematic transitions in aqueous suspensions of natural nontronite clay. Influence of particle anisotropy. 1. Features of the I/N transition. Langmuir 2008, 24, 3127–3139. [Google Scholar] [CrossRef]
  105. Lou, X.; Daussin, R.; Cuenot, S.; Duwez, A.S.; Pagnoulle, C.; Detrembleur, C.; Bailly, C.; Jérôme, R. Synthesis of pyrene-containing polymers and noncovalent sidewall functionalization of multiwalled carbon nanotubes. Chem. Mater. 2004, 16, 4005–4011. [Google Scholar] [CrossRef]
  106. Bahun, G.J.; Wang, C.; Adronov, A. Solubilizing single-walled carbon nanotubes with pyrene-functionalized block copolymers. J. Polym. Sci. Part A Polym. Chem. 2006, 44, 1941–1951. [Google Scholar] [CrossRef]
  107. Meuer, S.; Braun, L.; Zentel, R. Solubilisation of multi walled carbon nanotubes by α-pyrene functionalised PMMA and their liquid crystalline self-organisation. Chem. Commun. 2008, 27, 3166. [Google Scholar] [CrossRef] [PubMed]
  108. Beek, W.J.E.; Wienk, M.M.; Janssen, R.A.J. Efficient hybrid solar cells from zinc oxide nanoparticles and a conjugated polymer. Adv. Mater. 2004, 16, 1009–1013. [Google Scholar] [CrossRef]
  109. Grätzel, M. Solar energy conversion by dye-sensitized photovoltaic cells. Inorg. Chem. 2005, 44, 6841–6851. [Google Scholar] [CrossRef]
  110. Suri, P.; Mehra, R.M. Effect of electrolytes on the photovoltaic performance of a hybrid dye sensitized ZnO solar cell. Sol. Energy Mater. Sol. Cells 2007, 91, 518–524. [Google Scholar] [CrossRef]
  111. Sites, J.; Pan, J. Strategies to increase CdTe solar-cell voltage. Thin Solid Film. 2007, 515, 6099–6102. [Google Scholar] [CrossRef]
  112. Huynh, W.U.; Dittmer, J.J.; Alivisatos, A.P. Hybrid nanorod-polymer solar cells. Science 2002, 295, 2425–2428. [Google Scholar] [CrossRef] [Green Version]
  113. Zorn, M.; Zentel, R. Liquid crystalline orientation of semiconducting nanorods in a semiconducting matrix. Macromol. Rapid Commun. 2008, 29, 922–927. [Google Scholar] [CrossRef]
  114. Zorn, M.; Tahir, M.N.; Bergmann, B.; Tremel, W.; Grigoriadis, C.; Floudas, G.; Zentel, R. Orientation and dynamics of ZnO nanorod liquid crystals in electric fields. Macromol. Rapid Commun. 2010, 31, 1101–1107. [Google Scholar] [CrossRef]
  115. Zorn, M.; Weber, S.A.L.; Tahir, M.N.; Tremel, W.; Butt, H.J.; Berger, R.; Zentel, R. Light induced charging of polymer functionalized nanorods. Nano Lett. 2010, 10, 2812–2816. [Google Scholar] [CrossRef] [PubMed]
  116. Zur Borg, L.; Domanski, A.L.; Berger, R.; Zentel, R. Photoinduced Charge separation of self-organized semiconducting superstructures composed of a functional polymer-TiO2 hybrid. Macromol. Chem. Phys. 2013, 214, 975–984. [Google Scholar] [CrossRef]
  117. Zur, B.L.; Domanski, A.L.; Breivogel, A.; Bürger, M.; Berger, R.; Heinze, K.; Zentel, R. Light-induced charge separation in a donor–chromophore–acceptor nanocomposite Poly[TPA-Ru(Tpy)2]@ZnO. J. Mater. Chem. C 2013, 1, 1223–1230. [Google Scholar] [CrossRef]
  118. Oschmann, B.; Bresser, D.; Tahir, M.N.; Fischer, K.; Tremel, W.; Passerini, S.; Zentel, R. Polyacrylonitrile block copolymers for the preparation of a thin carbon coating around TiO2 nanorods for advanced lithium-ion batteries. Macromol. Rapid Commun. 2013, 34, 1693–1700. [Google Scholar] [CrossRef] [PubMed]
  119. Oschmann, B.; Tahir, M.N.; Mueller, F.; Bresser, D.; Lieberwirth, I.; Tremel, W.; Passerini, S.; Zentel, R. Precursor polymers for the carbon coating of Au@ZnO multipods for application as active material in lithium-ion batteries. Macromol. Rapid Commun. 2015, 36, 1075–1082. [Google Scholar] [CrossRef] [PubMed]
  120. Tahir, M.N.; Oschmann, B.; Buchholz, D.; Dou, X.; Lieberwirth, I.; Panthöfer, M.; Tremel, W.; Zentel, R.; Passerini, S. Extraordinary performance of carbon-coated anatase TiO2 as sodium-ion anode. Adv. Energy Mater. 2016, 6, 1–9. [Google Scholar] [CrossRef] [Green Version]
  121. Xu, W.; Ji, W.; Jing, P.; Yuan, X.; Wang, Y.A.; Xiang, W.; Zhao, J. Efficient inverted quantum-dot light-emitting devices with TiO2/ZnO bilayer as the electron contact layer. Opt. Lett. 2014, 39, 426. [Google Scholar] [CrossRef]
  122. Dai, X.; Zhang, Z.; Jin, Y.; Niu, Y.; Cao, H.; Liang, X.; Chen, L.; Wang, J.; Peng, X. Solution-processed, high-performance light-emitting diodes based on quantum dots. Nature 2014, 515, 96–99. [Google Scholar] [CrossRef]
  123. Liu, Y.-Q.; Zhang, D.-D.; Wei, H.-X.; Ou, Q.-D.; Li, Y.-Q.; Tang, J.-X. Highly efficient quantum-dot light emitting diodes with sol-gel ZnO electron contact. Opt. Mater. Express 2017, 7, 2161. [Google Scholar] [CrossRef]
  124. Talapin, D.V.; Mekis, I.; Götzinger, S.; Kornowski, A.; Benson, O.; Weller, H. CdSe/CdS/ZnS and CdSe/ZnSe/ZnS core-shell-shell nanocrystals. J. Phys. Chem. B 2004, 108, 18826–18831. [Google Scholar] [CrossRef]
  125. Kwak, J.; Bae, W.K.; Lee, D.; Park, I.; Lim, J.; Park, M.; Cho, H.; Woo, H.; Yoon, D.Y.; Char, K.; et al. Bright and efficient full-color colloidal quantum dot light-emitting diodes using an inverted device structure. Nano Lett. 2012, 12, 2362–2366. [Google Scholar] [CrossRef] [PubMed]
  126. Shirasaki, Y.; Supran, G.J.; Tisdale, W.A.; Bulović, V. Origin of efficiency roll-off in colloidal quantum-dot light-emitting diodes. Phys. Rev. Lett. 2013, 110, 1–5. [Google Scholar] [CrossRef] [Green Version]
  127. Lim, J.; Jeong, B.G.; Park, M.; Kim, J.K.; Pietryga, J.M.; Park, Y.S.; Klimov, V.I.; Lee, C.; Lee, D.C.; Bae, W.K. Influence of shell thickness on the performance of light-emitting devices based on CdSe/Zn1−xCdxS core/shell heterostructured quantum dots. Adv. Mater. 2014, 26, 8034–8040. [Google Scholar] [CrossRef]
  128. Kim, H.H.; Park, S.; Yi, Y.; Son, D.I.; Park, C.; Hwang, D.K.; Choi, W.K. Inverted quantum dot light emitting diodes using polyethylenimine ethoxylated modified ZnO. Sci. Rep. 2015, 5, 1–5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Shen, H.; Cao, W.; Shewmon, N.T.; Yang, C.; Li, L.S.; Xue, J. High-efficiency, low turn-on voltage blue-violet quantum-dot-based light-emitting diodes. Nano Lett. 2015, 15, 1211–1216. [Google Scholar] [CrossRef] [PubMed]
  130. Pan, J.; Chen, J.; Zhao, D.; Huang, Q.; Khan, Q.; Liu, X.; Tao, Z.; Zhang, Z.; Lei, W. Surface plasmon-enhanced quantum dot light-emitting diodes by incorporating gold nanoparticles. Opt. Express 2016, 24, A33. [Google Scholar] [CrossRef]
  131. Jung, H.; Chung, W.; Lee, C.H.; Kim, S.H. Fabrication of white light-emitting diodes based on UV light-emitting diodes with conjugated polymers-(CdSe/ZnS) quantum dots as hybrid phosphors. J. Nanosci. Nanotechnol. 2012, 12, 5407–5411. [Google Scholar] [CrossRef]
  132. Park, J.S.; Han, J.; Ha, J.S.; Seong, T.Y. Polarity dependence of the electrical characteristics of Ag reflectors for high-power GaN-based light emitting diodes. Appl. Phys. Lett. 2014, 104, 1–5. [Google Scholar] [CrossRef]
  133. Li, Q.; Zhang, W.C.; Wang, C.F.; Chen, S. In situ access to fluorescent dual-component polymers towards optoelectronic devices via inhomogeneous biphase frontal polymerization. RSC Adv. 2015, 5, 102294–102299. [Google Scholar] [CrossRef]
  134. Zorn, M.; Ki Bae, W.; Kwak, J.; Lee, H.; Lee, C.; Zentel, R.; Char, K. Quantum dot-block copolymer hybrids with improved properties and their application to quantum dot light-emitting devices. ACS Nano 2009, 3, 1063–1068. [Google Scholar] [CrossRef]
  135. Kwak, J.; Bae, W.K.; Zorn, M.; Woo, H.; Yoon, H.; Lim, J.; Kang, S.W.; Weber, S.; Butt, H.-J.; Zentel, R.; et al. Characterization of quantum dot/conducting polymer hybrid films and their application to light-emitting diodes. Adv. Mater. 2009, 21, 5022–5026. [Google Scholar] [CrossRef] [PubMed]
  136. Menk, F.; Fokina, A.; Oschmann, B.; Bauer, T.; Nyquist, Y.; Braun, L.; Kiehl, J.; Zentel, R. Functionalization of P3HT with various mono- and multidentate anchor groups. J. Braz. Chem. Soc. 2017. [Google Scholar] [CrossRef]
  137. zur Borg, L.; Lee, D.; Lim, J.; Bae, W.K.; Park, M.; Lee, S.; Lee, C.; Char, K.; Zentel, R. The effect of band gap alignment on the hole transport from semiconducting block copolymers to quantum dots. J. Mater. Chem. C 2013, 1, 1722. [Google Scholar] [CrossRef]
  138. Fokina, A.; Lee, Y.; Chang, J.H.; Braun, L.; Bae, W.K.; Char, K.; Lee, C.; Zentel, R. Side-chain conjugated polymers for use in the active layers of hybrid semiconducting polymer/quantum dot light emitting diodes. Polym. Chem. 2016, 7, 101–112. [Google Scholar] [CrossRef] [Green Version]
  139. Kim, W.D.; Kim, D.; Yoon, D.-E.; Lee, H.; Lim, J.; Bae, W.K.; Lee, D.C. Pushing the efficiency envelope for semiconductor nanocrystal-based electroluminescence devices using anisotropic nanocrystals. Chem. Mater. 2019, 31, 3066–3082. [Google Scholar] [CrossRef]
  140. Cheng, T.; Wang, Z.; Jin, S.; Wang, F.; Bai, Y.; Feng, H.; You, B.; Li, Y.; Hayat, T.; Tan, Z. Blue LEDs: Pure blue and highly luminescent quantum-dot light-emitting diodes with enhanced electron injection and exciton confinement via partially oxidized aluminum cathode (advanced optical materials 11/2017). Adv. Opt. Mater. 2017, 5. [Google Scholar] [CrossRef]
  141. Fu, Y.; Jiang, W.; Kim, D.; Lee, W.; Chae, H. Highly efficient and fully solution-processed inverted light-emitting diodes with charge control interlayers. ACS Appl. Mater. Interfaces 2018, 10, 17295–17300. [Google Scholar] [CrossRef]
  142. Khodabakhshi, E.; Klöckner, B.; Zentel, R.; Michels, J.J.; Blom, P.W.M. Suppression of electron trapping by quantum dot emitters using a grafted polystyrene shell. Mater. Horizons 2019, 6, 2024–2031. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic drawing of spherical and shape-anisotropic nanoparticles (left) and ways of stabilization by ionic and steric stabilization [8,10,20,21].
Figure 1. Schematic drawing of spherical and shape-anisotropic nanoparticles (left) and ways of stabilization by ionic and steric stabilization [8,10,20,21].
Inorganics 08 00020 g001
Figure 2. The coating of nanoparticles can either be done by a grafting from (a) or a grafting to (b) process. In the last case the ligand exchange is shown; reproduced with permission from [24]; published by Wiley-VCH, 2015.
Figure 2. The coating of nanoparticles can either be done by a grafting from (a) or a grafting to (b) process. In the last case the ligand exchange is shown; reproduced with permission from [24]; published by Wiley-VCH, 2015.
Inorganics 08 00020 g002
Figure 3. Recently established synthetic route [20,21] to blockcopolymers with multidentate block for the ligand exchange on nanoparticles.
Figure 3. Recently established synthetic route [20,21] to blockcopolymers with multidentate block for the ligand exchange on nanoparticles.
Inorganics 08 00020 g003
Figure 4. Visualization (A): upper line, AFM) and characterization by light scattering (B): lower line) of properly coated, individually dispersed inorganic nanoparticles (see [39] for more information); reproduced with permission of the ACS [39], published by the ACS, 2008.
Figure 4. Visualization (A): upper line, AFM) and characterization by light scattering (B): lower line) of properly coated, individually dispersed inorganic nanoparticles (see [39] for more information); reproduced with permission of the ACS [39], published by the ACS, 2008.
Inorganics 08 00020 g004
Figure 5. Ways to destabilize polymer coated nanoparticles with an external stimulus and to induce a coagulation. (a) Nanoparticles coated with a polymer with an lower critical solution temperature coagulate during temperature increase [39]; (b) Nanoparticles coated with polymers with a photosensitive group with in the polymer backbone allow it to split-off the stabilizing polymer corona [81]; lower part reproduced with permission of [81], published by Wiley-VCH, 2014.
Figure 5. Ways to destabilize polymer coated nanoparticles with an external stimulus and to induce a coagulation. (a) Nanoparticles coated with a polymer with an lower critical solution temperature coagulate during temperature increase [39]; (b) Nanoparticles coated with polymers with a photosensitive group with in the polymer backbone allow it to split-off the stabilizing polymer corona [81]; lower part reproduced with permission of [81], published by Wiley-VCH, 2014.
Inorganics 08 00020 g005
Figure 6. Liquid crystalline phases formed from shape anisotropic inorganic nanoparticles in highly concentrated solution [20]; (a) polarizing microscopy showing smectic and nematic textures with increasing temperature. (b) SEM microscopy of the dried phase consisting of TiO2 nano-rods; reproduced with permission of [20], published by Wiley-VCH, 2007.
Figure 6. Liquid crystalline phases formed from shape anisotropic inorganic nanoparticles in highly concentrated solution [20]; (a) polarizing microscopy showing smectic and nematic textures with increasing temperature. (b) SEM microscopy of the dried phase consisting of TiO2 nano-rods; reproduced with permission of [20], published by Wiley-VCH, 2007.
Inorganics 08 00020 g006
Figure 7. Schematic representation of a photo-voltaic cell optimized for an effective transport of the negative charge to the electrode and an assembly of ZnO nano-rods on an electrode; lower part reproduced with permission of [114], published by Wiley-VCH, 2010.
Figure 7. Schematic representation of a photo-voltaic cell optimized for an effective transport of the negative charge to the electrode and an assembly of ZnO nano-rods on an electrode; lower part reproduced with permission of [114], published by Wiley-VCH, 2010.
Inorganics 08 00020 g007
Figure 8. Schematic presentation of Kelvin-probe AFM measurements (left), which allow a determination of charging of nanoobjects and measurements (right), which prove the transfer of electrons from the coated semiconducting polymer to the TiO2 nanoparticles (see [115,116,117]); right part reproduced with permission of [115], published by the ACS, 2010.
Figure 8. Schematic presentation of Kelvin-probe AFM measurements (left), which allow a determination of charging of nanoobjects and measurements (right), which prove the transfer of electrons from the coated semiconducting polymer to the TiO2 nanoparticles (see [115,116,117]); right part reproduced with permission of [115], published by the ACS, 2010.
Inorganics 08 00020 g008
Figure 9. Coating of TiO2 nano-rods with block copolymers with a graphitable polyacrylonitrile block and their conversion into a nanocomposite with nano-rods embedded in graphite [118] (a) and the use of this nanocomposite as anode (b); reproduced with permission of [118], published by Wiley-VCH, 2013.
Figure 9. Coating of TiO2 nano-rods with block copolymers with a graphitable polyacrylonitrile block and their conversion into a nanocomposite with nano-rods embedded in graphite [118] (a) and the use of this nanocomposite as anode (b); reproduced with permission of [118], published by Wiley-VCH, 2013.
Inorganics 08 00020 g009
Figure 10. Block copolymers based on polyphenylene-vinylene [116] (A), polythiophene [136] (B) and polytriarylamine v [134] (C) and an anchor block for oxidic or sulfidic quantum dots.
Figure 10. Block copolymers based on polyphenylene-vinylene [116] (A), polythiophene [136] (B) and polytriarylamine v [134] (C) and an anchor block for oxidic or sulfidic quantum dots.
Inorganics 08 00020 g010
Figure 11. CdSe QDs as light emitting material in a matrix of semiconducting polymer for the preparation of an QLED; (a) dispersibility as viewed from the top and the side [135] (note the homogeneous dispersion of the polymer coated hybrids); (b) electroluminescence from 3 hybrid systems (QH1 to QH3, see ref. for assignment); they are identical to the electroluminescence of the pure QDs; reproduced with permission of [135], published by Wiley-VCH, 2009.
Figure 11. CdSe QDs as light emitting material in a matrix of semiconducting polymer for the preparation of an QLED; (a) dispersibility as viewed from the top and the side [135] (note the homogeneous dispersion of the polymer coated hybrids); (b) electroluminescence from 3 hybrid systems (QH1 to QH3, see ref. for assignment); they are identical to the electroluminescence of the pure QDs; reproduced with permission of [135], published by Wiley-VCH, 2009.
Inorganics 08 00020 g011
Figure 12. Scheme of the band gap alignment in an QLED with inverted device architecture (left) as well as 3 new monomers for the preparation of a semiconducting blockcopolymer; see [138] for more details.
Figure 12. Scheme of the band gap alignment in an QLED with inverted device architecture (left) as well as 3 new monomers for the preparation of a semiconducting blockcopolymer; see [138] for more details.
Inorganics 08 00020 g012
Figure 13. Results of the comparison of CdSe QLEDs with a single layer of fluorescent QDs (left, reference) as well as polymer coated hybrids and blends of unfunctionalized QDs and polymer; see [25] for more details; reproduced with permission of [25], published by Wiley-VCH, 2016.
Figure 13. Results of the comparison of CdSe QLEDs with a single layer of fluorescent QDs (left, reference) as well as polymer coated hybrids and blends of unfunctionalized QDs and polymer; see [25] for more details; reproduced with permission of [25], published by Wiley-VCH, 2016.
Inorganics 08 00020 g013
Figure 14. Structure of InP QDs coated with a ZnSeS shell (a) as well as the electroluminescence of QLEDs (b) prepared from such QDs after coating with a semiconducting polymer (a block copolymer from M2 in Figure 12).
Figure 14. Structure of InP QDs coated with a ZnSeS shell (a) as well as the electroluminescence of QLEDs (b) prepared from such QDs after coating with a semiconducting polymer (a block copolymer from M2 in Figure 12).
Inorganics 08 00020 g014
Figure 15. Schematic representation (a) of the general concept of preventing traps for charges by placing fluorescent guests inside a semiconducting polymer; coating the QD with a thin polystyrene shell (see energy diagram in (b)) can block the transfer of charges to the QD; the transfer of energy from the semiconducting matrix to the QD can, however, still proceed via Foerster transfer. The resulting electroluminescence (c) is voltage independent. See [142] for more information.
Figure 15. Schematic representation (a) of the general concept of preventing traps for charges by placing fluorescent guests inside a semiconducting polymer; coating the QD with a thin polystyrene shell (see energy diagram in (b)) can block the transfer of charges to the QD; the transfer of energy from the semiconducting matrix to the QD can, however, still proceed via Foerster transfer. The resulting electroluminescence (c) is voltage independent. See [142] for more information.
Inorganics 08 00020 g015

Share and Cite

MDPI and ACS Style

Zentel, R. Polymer Coated Semiconducting Nanoparticles for Hybrid Materials. Inorganics 2020, 8, 20. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8030020

AMA Style

Zentel R. Polymer Coated Semiconducting Nanoparticles for Hybrid Materials. Inorganics. 2020; 8(3):20. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8030020

Chicago/Turabian Style

Zentel, Rudolf. 2020. "Polymer Coated Semiconducting Nanoparticles for Hybrid Materials" Inorganics 8, no. 3: 20. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8030020

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

Article Metrics

Back to TopTop