Jalal Azadmanjiri
*a,
Christopher C. Berndt
ab,
James Wang
a,
Ajay Kapoor
a and
Vijay K. Srivastavac
aSchool of Engineering, Faculty of Science, Engineering and Technology, Swinburne University of Technology, Hawthorn, Victoria 3122, Australia. E-mail: jazadmanjiri@swin.edu.au; Fax: +61 3 9214 8264; Tel: +61 3 9214 5913
bDepartment of Materials Science and Engineering, Stony Brook University, Stony Brook, NY 11794, USA
cDepartment of Mechanical Engineering, Indian Institute of Technology, BHU, Varanasi-221005, India
First published on 8th November 2016
This article is a review on the nanolaminate composite materials from a materials science perspective. In fact, nanolaminate composite materials are a category of two-dimensional nanomaterials in the fields of chemistry, materials science and condensed matter physics. Miniaturization of materials to the nanometer scale is improving remarkably in science and technology. Nanolaminates constitute a unique class of nanomaterials that illustrate fascinating mechanical, physical, chemical and electrical properties and these attributes are increasingly being prospected for promising applications. Nanolaminates can typically be developed using bottom-up techniques that are designed with various stacking series along with layer thicknesses. The particular properties of fabricated nanolaminates can be determined by their arrangements, compositions and thicknesses. The properties can be established during the fabrication process by size control of each coating layer and interfacial chemical reactions between layers. Hence, fundamental understanding of nanoscale layered-engineering during the creation of a nanolaminate structure enables tailoring of the overall performance of these composite materials. This work demonstrates the bottom-up physical and chemical approaches for the fabrication of nanolaminates. The influence of the interface layer in the nanolaminate composite materials is considered from the viewpoint of conferring high performance characteristics. The desirable physical attributes will be pursued by incorporation of dopants and site-engineering techniques on various materials that would improve the nanolaminate properties. The final section concludes with an outlook on future directions in this technological field.
2D nanosheets and nanolaminates that are absolutely dense and ultra-fine grained could be produced for engineering applications to bring benefit of enhanced mechanical properties for devices such as energy storage. Nanolaminates can be fabricated using bottom-up8–12 deposition and in some cases by top-down13 techniques that are designed with various stacking series and layer thicknesses. The properties of created nanolaminates be determined by their compositions, arrangements and thicknesses. These can be demonstrated within the synthesis process through thickness control of each layer and interfacial chemical reaction between layers.
The rapid advances in nanoscience and technology provide new opportunities in achieving highly efficient nanolaminates. Particularly, physical properties such as a significant surface area and high efficient interface structure markedly improve the efficiency of such structures. Therefore, nanolaminated structures are increasingly important for the growing of innovative devices and have attracted excellent interest. Towards this aim, the present review article is focused on the developments in the structure, interface role and applications of nanolaminates; with an emphasis on recently published works.
Bottom-up | Building block | Applications | Reference |
---|---|---|---|
Self-assembly | p-type oligo (p-phenylenevinylene)s and n-type perylenebisimides thin films | Photoelectronic applications | 22 |
Self-assembly | Gold–polypyrrole amphiphiles | Electrocatalysis, chemical sensors and microelectronic devices | 22 |
Langmuir–Blodgett | Biomimetic lamellar lipid | Biological membranes | 26 |
Langmuir–Blodgett | Poly(3-hexyl thiophene) film | Glucose biosensor | 27 |
Layer-by-layer | Graphene/polyaniline (PANI) hybrid electrode | Supercapacitors | 28 |
Layer-by-layer | Graphene/palladium (Pd) nanoparticles films | Sensors | 28 |
Atomic layer deposition | Metal oxides (TiO2, Al2O3, etc.) thin films | Dielectric layers, insulating layer, anti-reflective layers, etc. | 29 |
Atomic layer deposition | Poly(methyl methacrylate) (PMMA), polyethylene, etc. | Polymer surface functionalization, creation of composites, diffusion barriers, etc. | 30 |
Electrochemical deposition | Platinum on alloys (such as Co-super alloy, Inconel, pure Al, Al–Ti alloy, graphite) | Batteries, fuel cells and capacitors | 31 |
Chemical vapor deposition | Gallium arsenide (GaAs) | Electronic (such as, ICs) and photovoltaic devices | 32 |
Chemical vapor deposition | Tungsten carbide | Wear resistance | 33 |
Magnetron sputtering | Nd-doped SnO2 thin films | Solar sell devices | 34 |
Magnetron sputtering | Al/CuO | MEMS energy sources, micro-initiators | 35 |
PLD is a fast technique (e.g., 1.3 and 2.0 monolayers for iron and silicon deposition per minute, respectively, using a Nd:YAG laser44) and the stoichiometry of the deposited film is very close to that of the target material; hence, it is possible to fabricate stoichiometric thin films from a single alloy bulk target.45 The properties of the fabricated thin film can also be controlled in terms of crystalline structure, thickness and composition.45 PLD does not need expensive or corrosive precursors, and large volume targets are not necessary with this technique compared to the other thin film methods such as magnetron sputtering. However, clusters and droplets may be observed at the surface of the deposited films that lead to higher roughness and this is a main drawback for optical, and electrical, etc.46 applications of the deposited film.
The microstructure and properties of the nanolaminates depend on their composition and the total nanolaminated thickness depends on the thickness of the individual layers. The structure and composition can be manipulated by adjusting the process parameters. Sputtered thin films manufactured by magnetron sputtering have excellent chemical and microstructural uniformity. As well, almost any metal target material can be sputtered without decomposition. The water-cooled target minimizes the effects of radiated heat so that thermal microstructural effects are diminished. However, slow deposition speed, low adhesion, and low density are several drawbacks for magnetron sputtering compared to other technologies.
CVD may be implemented in different equipment formats; however, most CVD processes may be classified as either low pressure CVD (LPCVD) or ultra-high vacuum CVD (UHCVD) processes. LPCVD is carried out under sub-atmospheric pressures to prevent unwanted reactions and confer a more uniform coating thickness. UHCVD is performed under low atmospheric pressures (∼10−6 Pa). The CVD precursor gases that react with the substrate must be highly volatile, otherwise it is difficult to deliver them to the reaction chamber. The gaseous by-products of the CVD process are usually toxic, which is a disadvantage that can be mitigated by appropriate design of the CVD reaction chamber and post-process disposal that employs gas cleaning technology.
Fig. 5 demonstrates direct deposition of graphene onto a quartz dielectric substrate, which offers superior electronic properties, by a single-step CVD process.10 The graphene layer is formed using surface catalytic decomposition of hydrocarbon precursors on thin copper films pre-coated onto a quartz substrate.10 The copper films evaporated during or instantly after graphene growth, thereby resulting in a graphene layer that is coated directly onto the quartz substrate.
Fig. 5 A schematic for demonstration of CVD process for deposition of graphene on quartz substrate. (a) A thin film of copper is deposited on the dielectric surface, firstly, (b) during the CVD process, (c) the copper thin film evaporates and (d) leaving the graphene layer on the quartz substrate. Reproduced with permission from data published in ref. 10 Copyright 2010, American Chemical Society. |
Fig. 7 A schematic illustration of an ALD method where the thin film layers cyclically grow for a binary compound from gaseous precursors. |
The ALD method has been widely considered by researchers in nanotechnology areas due to its control over the deposited layer thickness of materials and the potential for modification of chemical and physical properties in the nanoscale range.11 There is a comprehensive review on metal and nitride thin films fabricated by ALD for semiconductor device applications by Kim et al.48 The surface chemistry of the trimethylaluminum [Al(CH3)3] (TMA) and H2O process by ALD is elaborated by Puurunen et al.49 Moreover, HfO2 is considered for electronic applications50 and the electrical and mechanical properties of some metal oxides such as ZnO,51 TiO252 and Al2O353 have been investigated systematically. Other groups have developed deposition processes for metals such as platinum and ruthenium for electrode material applications.49 Slowness of deposition is the major limitation of the ALD technique; usually a fraction of coverage is deposited in each cycle. However, a very thin layer is required for next-generation electronic devices and thus the slowness of ALD method may not a critical issue.
Fig. 8 A schematic demonstrating LB technique for producing mono- and multimolecular films and 3 types of LB deposited layers on a hydrophobic surface. |
These methods have some disadvantages. LB is an expensive and time consuming method that requires specialized instrumentation. It also requires specific and limited molecules (e.g., amphiphilic molecules) to prepare the films.54 Additionally, a weak molecular interaction between the film and the solid support is a disadvantage for LB.54 Deposited films using the SAM process also have limited stability and strength under ambient and physiological conditions.
The LbL method generally occurs through a variety of deposition techniques such as dip-coating, spin-coating, spraying, and perfusion that have been extensively reviewed in previous work.57–60 The architecture and properties of the deposited films by the LbL method can be well-controlled at the nanometer-scale level.55 There are several molecular interactions between thin film layers for the LbL assembly technique that have been overviewed by J. Borges and J.F. Mano.55 Fig. 10 shows a schematic for a LbL self-assembly of a multilayer coating by sequential adsorption of oppositely charged polyelectrolytes.
Fig. 11 exhibits a schematic of some typical ultrathin 2D nanomaterials, such as transition metal dichalcogenides (TMDs), hexagonal boron nitride (h-BN), covalent organic frameworks (COFs), metal–organic frameworks (MOFs), layered double hydroxides (LDHs), black phosphorus (BP), MXenes, metals and oxides.
Fig. 11 A schematic demonstration of various kinds of the typical ultrathin 2D nanomaterials, such as LDHs, h-BN, metals, oxides, BP, TMDs, Mxenes, COFs and MOFs. Reproduced from data published in ref. 65 Copyright 2015, American Chemical Society. |
Nanolaminates are a subclass of ultrathin 2D nanomaterials that incorporate binary and ternary elements of various layer thicknesses, Fig. 12. They are atomic size sandwiches made with simply 2 or even up to 200000 layers. The nanolaminate structures are fully dense, ultra-fine grained solids that provide an increased concentration of interface defects. The properties of the nanolaminates depend on their particular arrangements, compositions, thicknesses and are always different from those of the bulk form components. The different properties arise due to the nanometer-scale environment of the atoms in each layer. Nanolayers with binary and ternary compounds, as well as atoms within nanolayers that are powerfully affected by the interfaces between the component sublayers, will be described in the following section.
Binary metal oxides can be divided in two categories of dielectric binary oxides (e.g.; MgO, Al2O3, SiO2, TiO2, ZrO2, HfO2, V2O5, Nb2O5, Ta2O5, Sc2O3, Y2O3, La2O3, CeO2, Nd2O3, Er2O3),68–71 and conductors/semiconductors of binary oxides (e.g.; WO3, MnOx, Co3O4, NiO, ZnO, Ga2O3, In2O3, SnO2, Sb2O3).68,72,73
It has been shown that the nanolaminate properties of several binary oxides can be merged advantageously.67 Hard coatings, optical filters and X-ray mirrors are examples that exploit binary metal oxide multilayers.74–76 In addition to good chemical stability, excellent mechanical and optical properties; advanced dielectric properties can be achieved with multilayered structures. As-deposited binary metal oxides usually show relatively large leakage currents and low breakdown resistance, which are essential for energy storage applications.
For example, Al2O3, which is an excellent insulator with high band gap (5.95 eV and κ = 9), may be combined with oxides of high dielectric constant (high-κ), like HfO2 (κ = 25), TiO2 (κ = 80) or ZrO2 (κ = 25).77–81 If the single binary layer thickness attains several nanometers, then the dielectric with a relatively high κ compared to that of the reference SiO2 (κ = 3.9) can be constructed without strong negative dependence on the voltage or temperature.67 Nevertheless, the quantities of dielectric constant that can be attained are maximized, resulting in values of up to 8, 20 and 26 in amorphous Al2O3/TiO2 nanolaminates,80 nanocrystalline Al2O3/TiO2 nanolaminates79 and Al2O3/TiO2/Al2O3 trilayers,82 respectively.
Li et al.83 indicated that modifying the interfacial composition of nanolaminate structures, including layers of two different materials (AlOx and TiOy) with low dielectric constant, can contribute to dielectric polarization under an externally applied field. Hence, the dielectric constant of the layered structures is enhanced, Fig. 13; which may be attributed to the internal barrier layer capacitance because of the attendance of electrical inhomogeneity in thin film materials. These results demonstrate that the dielectric properties of materials can be altered considerably by employing the design concept of nanolaminate structures.
Fig. 13 (a) A schematic of the Al2O3 and TiO2 nanolaminates, (b) titanium and aluminum elemental maps obtained from energy filtered TEM images, and (c) bright field TEM image of AlOx–TiOy nanolaminates. Adapted with permission from data published in ref. 83 Copyright 2011, AIP Publishing LLC. |
ZnO/Al2O3 nanolaminates fabricated by atomic layer deposition (ALD) is another nanolaminate with sublayers of ZnO (semiconductor binary oxide, refractive index at λ = 400 nm is ∼2.4 for ZnO with 50 nm) and Al2O3 (dielectric binary oxide, refractive index at λ = 400 nm is ∼1.74 for Al2O3 with 50 nm).84 The third-order optical nonlinearity of ZnO/Al2O3 nanolaminates can be enhanced by nanoscale engineering of the thin film structure (refractive index for ZnO and Al2O3 in Al2O3/ZnO with 2 (50/50 nm) at λ = 400 nm is: 2.5 and 1.79, respectively).84 The grain size of the crystalline ZnO film was controlled by varying the thickness of the ZnO layers within the nanolaminate in which thin (2 nm) amorphous Al2O3 layers impede ZnO crystal growth.84 Nanoscale engineering on the nanolaminates was altered to attain a strong optical “third-harmonic generation” from the optimized nanolaminate structure compared to a ZnO reference film of comparable thickness.84
MXenes are a new family of 2D transition metal carbides and/or nitrides that are produced by etching and extracting an “A” layer from MAX phases13 (Note: MAX phases in the form of ternary nanolaminated compounds are mentioned in the next section). Fig. 14 shows a schematic of the MXene phase. The etching process is performed by immersing the MAX phase in strong etching solutions, such as hydrofluoric acid (HF) or ammonium bifluoride (NH4HF2) at room temperature. MXenes with binary compounds could be produced with compositions of M2X, M3X2, and M4X3.13,85
Fig. 14 A schematic illustration of exfoliation of “A” out of MAX phase to form MXene. Reproduced with permission from data published in ref. 93 Copyright 2012, American Chemical Society. |
MXene monolayers are predicted to be metallic since they show a high electron density near the Fermi level.86–90 Hence, multilayer MXenes are electronically conductive with a conductivity similar to multilayer graphene. In contrast to graphene, MXenes show hydrophilic behavior that offers dispersion in aqueous solutions.13 Since MXenes are materials with layered solids and the bonding strength between the layers is weak, then intercalation of the guest molecules in MXenes is possible. Therefore, it has been shown that MXenes can easily be intercalated with a range of organic molecules and inorganic salts that not only enable synthesis of various intercalation compounds, but also lead to new applications for these materials.13 MXene phases are being investigated further; for example their synthesis by exploring the structure and surface termination of MXenes to define their chemical formulas and control chemical composition.13 MXenes materials could be used for a broad range of applications including electronic devices, gas sensors, photocatalysis, reinforcement for composites, and energy storage materials.13,91,92
Fig. 15 A schematic illustration of structure of the MAX phases and the corresponding MXenes. Adapted with permission from data published in ref. 13 Copyright 2013, John Wiley and Sons. |
Metals are generally recognized using getting thermally and electrically conductive, plastically deformable at room temperature, machinable, thermal shock resistant, etcetera. On the other hand, ceramics are usually defined by high elastic moduli, high oxidation and corrosion resistance, good mechanical properties at high temperature, and so on. Recently, investigators have fabricated ternary Mn+1AXn (or MAX) phases with a naturally nanolaminated structure and unique combination of properties.94,95,97,99,100 In fact, the MAX phases are attractive because of their significant amalgamation of chemical, physical, electrical and mechanical properties, which combines the characteristics of metals and ceramics.95 For instance, MAX phases are usually resistant to oxidation and corrosion as well as being elastically stiff. However, simultaneously they display high electrical and thermal conductivities and are machinable. These properties origin from an intrinsically nanolaminated crystal structure, with Mn+1Xn slabs intercalated with pure A-element layers.95
The research on MAX phases has advanced by using innovative fabrication techniques of thin films. Magnetron sputtering and arc deposition have been employed to synthesis single-crystal materials by epitaxial growth that have been further studied in terms of fundamental material properties.95 The required synthesis temperature for deposition of V2GeC and Cr2AlC using sputtering techniques is ∼450 °C101,102 whereas the deposition of MAX phases from the Ti–Si–C and Ti–Al–N systems usually requires ∼800 °C.95
First-principle calculations for predicting hypothetical MAX phases and their properties are in progress. New MAX phases with desired properties, such as Mo2Ga2C and Nb2GeC 97,100 have been discovered. Theoretical and experimental approaches are aimed at introducing dopants and forming solid solutions in a controlled manner. In this fashion, MAX phase solid solutions with attractive and novel properties may be identified in the new thin-film materials.
The role of interfaces between bilayers in materials with a nanolaminated structure is important and has been investigated in the fields of catalysis, corrosion science and inorganic semiconductor devices.103 However, the formation of interfacial layers in nanolaminates are not understood well because of an insufficient in situ characterization; thereby leading to poor control of the significant properties.104 Hence, these interfaces perform a critical role concerning the properties of these materials; which is in addition to the large lateral surface area that is a desirable characteristic for nanolaminates. For example, a chemical and/or physical interaction can occur at the interface103 when metal oxides and metals build an atomic-scale sandwich structure. The nanometer scale chemical and/or physical interaction will influence the chemical, physical or mechanical properties of the nanolaminates,14,103–110 Table 2.
Composition of the bilayers | Interfacial effect | Reference |
---|---|---|
Metal/molybdenum trioxide | Effective on chemical, electronic properties and work function | 103 |
Aluminum/copper(II) oxide | Controls the onset reaction temperature, reaction kinetic, and stability at low temperature | 104 and 106 |
Aluminum/copper/copper(II) oxide | Increases flame velocity and decreases reaction temperature | 105 |
Copper/zirconium | Enhances strength, tensile ductility and toughness | 14 |
Silicon nitride/titanium nitride | Enhances hardness and strength | 107 |
Copper/niobium | Enhances strength, tensile ductility and toughness | 108 |
Aluminum oxide/titanium dioxide/aluminum oxide | Promote the dielectric constant value | 109 |
Organic/clay | Ultralow thermal conductivity | 110 |
In an early effort, Kwon et al.104 employed a magnetron sputtering technique to fabricate high-quality aluminium/copper(II) oxide (Al/CuO) bilayers that were grown onto oxidized N type (100) double-side polished silicon, and studied chemical behavior of interfacial layer. Two sets of Al/CuO and Al/Al2O3/CuO (Al2O3, between Al and CuO) nanolaminates were fabricated to examine the interfacial chemistry and the role of exothermic reaction between these nanostructured materials. The synthesis of Al/CuO nanolaminates was implemented via magnetron sputtering on a piece of cleaned silicon wafer that was rinsed in deionized water and dried in a nitrogen gas steam. The substrate temperature was 10 °C during sputter deposition and the base pressure of the chamber was less than 10−5 Pa. CuO and Al layers were deposited using Cu and Al targets with purities >99.999%. Sputtering was carried out under an argon (Ar) and oxygen plasma at 400 W for Cu and only argon at 800 W for Al. The magnetron sputtering chamber was fully pumped out after each Al or CuO deposition cycle to avoid cross contamination and Al oxidation from residue oxygen during CuO deposition. The consecutive deposition of CuO and Al was performed without venting the chamber. The thicknesses of each CuO and Al layer were 100 nm and these were under less than 30 MPa stress.
High resolution transmission electron microscopy (HRTEM) on Al/CuO nanolaminates (Fig. 16) shows that an intermixed interfacial layer (a mixture of Cu, O and Al), which was irregular in thickness, is evident between the Al and CuO layers. The thickness of the interface layer is negligible in some regions and as thick as ∼5 nm in other regions. This varying structure arose due to the highly textured and 15 nm rough CuO surface.
Fig. 16 HRTEM images of the interface between sputter-deposited Al and CuO. Adapted with permission from data published in ref. 104 Copyright 2013, American Chemical Society. |
Further characterization of the intermixed interface layer region was executed on a 2 monolayer (ML) thin film of Al, which is equal to 0.5 nm, deposited onto a CuO surface. The thin layer was characterized via in situ XPS. The in situ XPS results (Fig. 17) show that the Cu 2p reduced from a higher binding energy of 933.5 eV (black pattern) to 932 eV (red pattern) after Al deposition.
Fig. 17 The (black) graph shows XPS Cu 2p spectra of CuO, the (blue) pattern is after 2 ALD cycles with trimethylaluminum (TMA, Al2(CH3)6) and water, and e-beam evaporated 2 monolayer of Al (red). The two dotted lines at 943.7 and 962 eV display the satellite peaks of Cu(II) of the underlying CuO. Adapted with permission from data published in ref. 104 Copyright 2013, American Chemical Society. |
These results illustrate that CuO [Cu(II)] is reduced to Cu2O [Cu(I)] in the interfacial region. These reduction reactions and the intermixed interface layer between Al and CuO are both effective with regard to the reduced onset reaction temperature, the reaction kinetics and stability at low temperature; all of which are important for reactive and energetic materials.
The mechanism of interface formation for Al/CuO nanolaminates, obtained from theoretical and experimental data by Kwon et al.,104 may be summarized as the deposition of isolated Al atoms on CuO surfaces leads to amorphization of the surface/interface through the insertion of Al atoms into substantial Cu sites (Fig. 18b). This mechanism results in a mixed interface composed of Al, Cu and O. The significant rearrangement of oxygen atoms around Al atoms confirms the experimental observation of CuO reduction. Fig. 18 depicts the Al/CuO reaction process, from the initial physical vapor deposition (Fig. 18a) to the formation of reduced products in the interface layer. The differential scanning calorimetric (DSC) thermal analysis data (Fig. 18c) supports this proposed mechanistic model.
Fig. 18 A schematic of (a) Al deposition on CuO, (b) Al/CuO reaction and amorphization process upon Al subsurface penetration and zoom around the Al site revealing the tetrahedral AlO4 structure and (c) DSC thermal analysis pattern of the Al/CuO bilayers without Al2O3 diffusion barrier. Reproduced with permission from data published in ref. 104 Copyright 2013, American Chemical Society. |
Kwon et al.104 in addition, fabricated Al/Al2O3/CuO nanolaminates to control the composition, thickness and conformity of the interface between the reactive materials of Al and CuO. Hence, an ultrathin Al2O3 interface was deposited between Al and CuO. Al2O3 was deposited using the ALD technique on CuO at 120 °C with alternating pulses of trimethylaluminum [TMA, Al2(CH3)6] for 2 seconds and heavy water (D2O) for 0.5 seconds. A separate N2 purge was applied for 10 minutes.115 The incoming Al species are CH3 decorated by implementation of the ALD method; hence the interface chemistry would be expected to differ from the previous deposition method that described the dual layer Al/CuO system.
Thermal analysis on the Al/Al2O3/CuO system displayed the presence of Al2O3 layers as thin as 0.5 nm that offered a barrier layer and which impact the long range diffusion of Al, Cu and O atoms; thereby enabling material combustion.104 The reactivity properties do not count on the Al2O3 thickness. Therefore, the composition and conformal coverage of the initial few Al2O3 monolayers on CuO are essential to confer thermal stability on this system. The TMA surface reactions with CuO in the Al/Al2O3/CuO system are the same as the previous dual Al/CuO system. TMA leads to the extraction of oxygen atoms from CuO and thus to reduction of CuO. Nevertheless, the presence of the surface methyl ligands (R–CH3) prevents Al diffusion into CuO.104 Fig. 19 diagrams the Al/Al2O3/CuO reaction process; from the beginning of ALD deposition to the final reduced products in the interface layer. The DSC data complement these results.
Fig. 19 A schematic of (a) Al2O3 deposition on CuO, (b) CuO reduction mechanism induced by the decomposition of TMA and (c) DSC thermal analysis pattern of the Al/CuO bilayers with Al2O3 diffusion barrier. Reproduced with permission from data published in ref. 104 Copyright 2013, American Chemical Society. |
In summary, there is a major exothermic reaction (at 530 °C) in Al/CuO nanolaminate with another minor exothermic reaction at lower temperatures (315 and 400 °C), while there are only two minor exothermic reactions (at 300 and 550 °C) for Al2O3 with 0.5 nm and (at 250 and 550 °C) for Al2O3 with 2.4 nm. The reactions arise because an interface layer is formed during physical deposition of Al. A mixture of Cu, O and Al is created through Al diffusion into CuO and this mixture of elements and a compound presence a low diffusion barrier. In contrast, deposition an Al2O3 layers, using TMA by the ALD method, produces a conformal coating that efficiently hampers Al diffusion; even for ultrathin layer thickness of ∼0.5 nm. This barrier results in better stability and durability at low temperature and reduced reactivity between Al and CuO.
In another report by Marín et al.105 the reactivity of thermite Al/CuO multilayered systems studied. Thermite multilayered systems are interesting structures as energetic layers on a chip that have potential applications such as micropropulsion in space, nanoairbags to drive fluids,105 MEMS energy sources,116 pressure-mediated molecular delivery117,118 and microinitiators.119,120 Multilayered structures are simple to control with regard to the thickness of each layer and the number of layers.105 These structural changes in a thermite multilayered system can be utilized to tune their energetic performance. Al/CuO was studied105 because the Al/CuO system has an energy density of 21 kJ cm−3, that is almost 3 times higher than trinitrotoluene (“TNT” at 7.6 kJ cm−3).106 The measured flame propagation velocity versus bilayer thickness in Al/CuO nanolaminates exhibit that the reaction-propagation velocity decreases from 90 to 2 m s−1 as the nanolaminate thickness increases from 50 to 2000 nm.105 If the bilayer thickness is decreased to below 25 nm, then the flame propagation velocity falls to zero. The flame velocity changes in the nanolaminates of less than 25 nm is due to spontaneous and uncontrolled intermixing at the interface during deposition (Fig. 20).
Fig. 20 A schematic of (a) Al/CuO nanolaminate with self-propagation rate and reaction heat of 44 m s−1 and 1 kJ g−1, respectively, (b) Al/Cu/CuO nanolaminate with improved self-propagation rate and reaction heat of 72 m s−1 and 1/3 kJ g−1, respectively. The heat flow curves also show that the reaction temperature is reduced for the Al/Cu/CuO nanolaminate and it accrued at two steps. The flame velocity of the both nanolaminates also is decreasing with increasing thickness of the bilayers (c). Reproduced with permission from data published in ref. 105 Copyright 2015, American Chemical Society. |
Marín's group has demonstrated that a thin layer of Cu of around 5 nm, deposited between each Al and CuO layer of an Al/CuO nanolaminates (i.e., the dual Al/CuO system changes to Al/Cu/CuO), alters the system reactivity. The flame velocity increases to 72 ± 1 m s−1 with the Al/Cu/CuO nanolaminate system, compared to 44 ± 1 m s−1 for the Al/CuO nanolaminates. The increase in flame temperature arises due to spontaneous diffusion of Cu into Al layers, which incubates the formation of an interfacial Al:Cu alloy with a melting temperature lower than pure Al.105 The newly-created alloy is responsible for the enhanced reactivity of the system.
Fig. 20 displays the enhanced reactivity of the Al/CuO nanolaminates. These results for Cu deposition between Al and CuO layers open new insights with regard to the interface engineering of nanoenergetic materials that incorporate a nanolaminate structure. The investigations such as Al/CuO and Al/Al2O3/CuO systems highlight the importance of the chemical characteristic of the diffusion barrier with regard to the interface structure of nanolaminates, also the crucial role of these interfacial layers concerning the behavior of nanostructured laminates. Thus, selective appropriate elements, synthesis process and precise characterization can be used to increase efficiency of reactive materials that are improving remarkably in science and technology.
Patscheider et al.107 sputtered several nanolayers of Si3N4, Si and Ti onto epitaxial TiN layers on MgO(001) to fabricate Si3N4/TiN(001), Si/TiN(001) and Ti/TiN(001). The hardness of the nanolaminates and the role of bilayer interfaces were investigated. The fabricated nanolaminates were characterized using in situ angle-resolved X-ray photoelectron (AR-XPS) spectroscopy. In situ AR-XPS allowed satellite intensity to be monitored; i.e., the positioning and local atomic configuration, which is expected to confer the physical properties of the nanolaminate ensemble. The satellite peak is collected from photoemission interaction of the X-ray with valence electrons of a sample. The satellite intensity is a direct measure of the interface valence electron density which is, in turn, linked to the interfacial bond strength and hardness.107 The thickness of the overlay materials (Si3N4, Ti and Si) on TiN was selected to be very thin to insure sufficient electron transparency so that the interfaces could be probed. Hence, only 4 monolayers (ML) were grown on the substrate [MgO(001)] within an ultrahigh vacuum system by using reactive magnetron sputter deposition.
The XPS spectra results indicate that both the width and intensity of Ti 2p satellites increased once TiN(001) was covered with 4 ML of Si3N4 at a substrate bias voltage (Vb) of −7 V (Fig. 21a). The development in intensity of Ti 2p satellites in Si3N4/TiN(001) is due to a high nitrogen (N) concentration around the surface and near-surface Ti atoms, thereby resulting in negative polarization because of the higher electronegativity of N than Ti.107 Electrons flow from the valence band of TiN to the overlayer material at the interface of Si3N4/TiN(001). Thus, electrostatic polarization along with covalent bonding is introduced that strengthens the Si3N4/TiN(001) interface.
Fig. 21 (a) XPS spectra of TiN(001) covered with 4 ML of Si3N4 at substrate bias voltage (Vb) of −7 V, (b) XPS spectra of Si3N4/TiN(001) at different Vb (−7, −150 and −250 V) as well as comparison of average bond coordination around Ti interface atoms at Vb of −7 V and −150 V. (c) XPS spectra of deposited Ti/TiN(001) and Si/TiN(001), and (d) demonstrate a model system for valance electron density at interface as well as band structure of Si3N4/TiN(001), Si/TiN(001) and Ti/TiN(001). Reproduced with permission from data published in ref. 107 Copyright 2011, American Physical Society. |
Si3N4/TiN(001) was also examined at different Vb (−7, −150 and −250 V). A distinct growth in the satellite intensity was observed when Vb increased (Fig. 21b). The increase of intensity of the Ti 2p satellites suggests that large interfacial electron accumulation and electrostatic polarization that is directly related to interfacial bonding and, hence, higher hardness could be obtained (Fig. 21b).107 On the other hand, XPS spectra from deposited Si/TiN(001) show that the intensities of the Ti 2p satellites are markedly reduced (Fig. 21c). Indeed, deposition of 4 ML of Ti leads to an almost complete loss of the satellite structure in Ti/TiN(001). Hence, based on the XPS intensities, it can be considered that the electron density in the interfaces of the fabricated nanolaminates are in this order: Si3N4/TiN(001) > Si/TiN(001) > Ti/TiN(001).
There is a direct relationship between satellite intensity and interfacial charge accumulation with band gap.107 Band gap values for Si3N4, Si and Ti can be found in Fig. 21. Materials with a high band gap offer larger interfacial charge accumulation.107 Fig. 21d demonstrates the interfacial band structure of Si3N4/TiN(001), Si/TiN(001) and Ti/TiN(001). A wide band gap in the thin film signifies that a larger number of free electrons are available to populate the interface valence band; hence, there is enhanced polarization, electron valence bonding and hardness. In contrast, when the contacting overlayer has a smaller band gap (Si), or none at all (Ti), electrons are donated by the overlayer coated to TiN and thereby reducing the interfacial polarization and interface strength.
It was found that satellite intensity, characterized by in situ AR-XPS, is a direct measure of the interface valence electron density. The latter has a direct relationship with interfacial bond strength and, thus, hardness. A Si3N4/TiN(001) nanolaminate with high N concentration at the interface, which accepts free electrons, can develop polarization of the interface. This effect is further enhanced with increasing the substrate bias voltage. Conversely, Si and Ti overlayers donate electrons to the TiN valence band and lead to charge depletion in the interfacial layers. Therefore, polarization at the interface of the Si3N4/TiN(001) nanolaminate increases the interfacial strength and provides enhanced hardness. On the other hand, low nitridation and polarization in Si/TiN(001) and Ti/TiN(001) results in an increasing contribution of Si and Ti atoms in contact with TiN and, thus, a weakened interface.
The mechanical properties of crystalline Cu/Cu–Zr glass nanolaminates were investigated.14 Single-phase bulk nanocrystalline Cu with an average grain size of 30 nm reveals a tensile strength of less than 800 MPa (Fig. 22a).106 In contrast, the tensile strength for the Cu/Zr nanolaminate with 5 nm Cu≈3Zr metallic glass thickness and a 35 nm nanocrystalline Cu layer is 1090 ± 20 MPa (Fig. 22a).14 The variance in tensile strengths illustrates that the amorphous interface layers in Cu/Zr nanolaminate have enhanced the strength compared to only nanocrystalline copper layers. In addition to this, the nanolaminate also exhibits a significant tensile ductility, with an elongation to failure of ε% = 13.8 ± 1.7%14 where ε is the strain.
Fig. 22 Microstructures and tensile properties of Cu/Zr nanolaminates. (a) Room-temperature tensile true stress–strain curves of the pure nanocrystalline (∼30 nm) of Cu, nanocrystalline–amorphous Cu/Zr nanolaminate and Cu/304 stainless steel crystalline multilayer with a sublayer thickness of 25 nm, all tested at the strain rate of 1 × 10−4 s−1. (b) TEM images from cross-sectional and top surface view of the as-deposited nanocrystalline Cu and amorphous Cu/Zr intermixing multilayer nanostructures. (c and d) MD simulations of Cu/Zr nanolaminates under periodic boundary conditions. Reproduced with permission from data published in ref. 14 Copyright 2007, National Academy of Sciences, USA. |
Other series of results swapped the amorphous Cu≈3Zr metallic glass layers with crystalline phase layers; e.g., 304 stainless steel (SS) or niobium (Nb), to make Cu/304 SS or Cu/Nb nanolaminates. The tensile strength increased to ∼1500 MPa for the Cu/304 SS nanolaminate (Fig. 22a). However, the elongation is <2%.123–125 A reduced ductility is the main limitation for all high strength crystalline/crystalline multilayer materials with nanometer scale size bilayers (bilayer ≤ 40 nm), and also in many single phase nanocrystalline metals such as nanocrystalline Cu (Fig. 22a).106,123–126 High-strength and high-ductility nanostructured Cu has been reported previously.127,128 Nevertheless, high tensile ductility in Cu/Zr nanolaminates with amorphous Cu≈3Zr is unexpected due to the near zero tensile ductility of bulk metallic glasses, and a typical low ductility in nanoscale microstructured materials.14
A further notable feature of a nanocrystalline–amorphous nanolaminate is toughness. The key to toughness is an appropriate blending of strength and ductility. Hence, high ductility in Cu/Zr nanolaminates would allow improved toughness due to significant plastic deformation (Fig. 22a), which is an uncommon behavior for metallic glasses. The room temperature behavior in Cu/Zr nanolaminate with a high flow stress of ∼1.09 GPa at a higher strain rate of 1 × 10−4 s−1 conflicts with the observations in pure nanocrystalline materials that indicated an alternative deformation mechanism.14 The plastic behavior observed in Cu/Zr nanolaminate is also in contrast with the Considère criterion that predicts the mathematical onset of necking when dσ/dε ≤ σ is reached at constant strain rate.14
In order to recognize the large tensile ductility and plastic behavior of the nanolaminates, the Cu/Zr nanolaminate was tested under HRTEM at different tensile strains.14 HRTEM shows that numerous dislocation storage or pileups was not revealed in the deformed nanocrystalline layers, except for a few deformation twins (Fig. 22b). Fig. 22b shows that one of the deformed twins tends to be terminated at the amorphous–crystalline interfaces (ACIs) or the Cu/Cu grain boundaries (GBs).14 Thus, in addition to GBs, the interface layer may have become a source for dislocation nucleation.14
Interaction between nanocrystalline layers and the nanoscale interface metallic glass in Cu/Zr nanolaminate may account for the large ductility enhancement in both phases.14 Wang et al. employed molecular dynamic (MD) simulations to validate this reasoning.14 The MD simulations display that dislocations are nucleated from ACIs or from GBs or intersections between ACIs and GBs; slipping across the nanocrystalline layer and being absorbed at the opposite ACI (Fig. 22c). The simulation results in addition illustrated that ACIs behave as inherent sinks of dislocations; that is, absorbing the dislocation value in the nanocrystalline Cu after plastic work is accomplished.129
The MD simulations indicate that the amorphous layers can impact dislocation structures created in the nanocrystalline layers (Fig. 22d). Many dislocations have been nucleated at the beginning of tensile deformation in the system, resulting in dense sessile dislocation forests (Fig. 22d). The dislocations form from intersections of two nonparallel stacking faults or twining systems.14 As the simulation time progresses, however, the dislocation density reduces significantly due to the attraction and annihilation of dislocations in the metallic glass layers (Fig. 22d).
Another method that often accompanies dislocation absorption is ACI sliding that spreads the dislocation core along the ACI.14,130 Therefore, according to experimental and simulation results, the nanoscale metallic glass layer not only retains large tensile plasticity itself, but plays the dominant mechanistic role; exhibiting the ability to behave as both a dislocation source and sink to intercede inelastic shear/slip transfer while avoiding extreme stress concentrations that would have led to fracture initiation.14
The extensive tensile ductility and nearly ideal plastic flow behavior in Cu/Zr nanolaminate propose that the nanocrystalline–metallic glass composite approach is an effective method towards developing materials with mechanical performance beyond those achievable from single phase elemental materials. The solid-state amorphization process that creates a crystalline–amorphous nanolaminate is evident in several material systems such as Cu/Ti, Ni/Ti, Cu/Hf, Ag/Zr, Ni/Zr, Ag/Hf, Ni/Nb, Ti/Si, Pd/Si and Cu/Nb.14,108
Fig. 23 Schematics for (a) a multilayered organic electronic device and a metal oxide shield layer, (b) layered structure of MoO3 oxidation states near an interface and plot of work function and conductivity versus MoO3 thickness for reactive and non-reactive interfaces between MoO3 and electrode materials. Reproduced with permission from data published in ref. 103 Copyright 2012, John Wiley and Sons. |
Greiner et al.103 studied the effect of the metal/metal oxide interface on the work function and electronic band structure. They examined MoO3 (purity 99.99%) films grown by vacuum sublimation on gold, nickel, molybdenum, vanadium and copper. MoO3 was selected since it is generally used as a hole-injection layer in organic devices to move an electron out of the device.132–135 Metals are used because they have different work functions and oxidation potentials.103
Mo cations have a nominal oxidation state of 6+ (Mo6+) in stoichiometric MoO3. However, the Mo cations of MoO3 are reduced to a lower oxidation state of Mo5+ and/or Mo4+ in the first few nanometers of each metal/MoO3 interface.103 The average oxidation state at each metal/MoO3 interface, work function and free energy of oxidation of the metals are summarized in Table 3.
Substrate material | Mo6+ | Mo5+ | Mo4+ | Work function (eV) | Free energy of oxidation (kJ mol−1) |
---|---|---|---|---|---|
Au | 60% | 40% | 0% | 5.3 | ∼−9 |
Ni | 50% | 50% | 0% | 5.0 | −419 |
Mo | 45% | 55% | 0% | 4.5 | −528 |
V | 7% | 36% | 57% | 4.2 | −800 |
Cu | 66% | 34% | 0% | 4.7 | −293 |
Molybdenum exhibits 6+ and 5+ oxidation states when MoO3 is in contact with the metals of Au, Ni, Mo and Cu; whereas Mo has 6+, 5+ and 4+ oxidation states when MoO3 is in contact with V. Copper is a unique electrode metal because it reacts with MoO3 to form a Cu–Mo–O alloy that incorporates both Mo6+ and Mo5+ valences.136,137
Oxidation-state and work-function profiles of each metal/MoO3 interface are shown in Fig. 24. These profiles display the relative Mo 3d peak areas for Mo6+, Mo5+ and Mo4+ peaks for each MoO3 film thickness, which were obtained from XPS experiments. Oxidation-state profiles show that the low-oxidation-state Mo cations persist further from the interface on the more reactive substrates than on the less-reactive substrates. The profiles also indicate that the d90% value is 6.4 nm for V, 6.0 nm for Mo, 33 nm for Ni and 2.8 nm for Au. Copper alloy (Cu–Mo–O) never achieves 90% (Note: d90% is assigned as the MoO3 thickness for which the Mo6+ peak is 90% of the total peak area).
Fig. 24 Profiles of (a) oxidation state and (b) work function for MoO3 films grown on various metal (Au, Ni, Mo, V and Cu) substrates. Adapted with permission from data published in ref. 103 Copyright 2012, John Wiley and Sons. |
The work-function depends on the MoO3 thickness (Fig. 24b).103 Fig. 24b indicates that work function tends to increase with distance far away from the metal/MoO3 interface and finally plateaus at a value of ∼6.85 to 6.9 eV (Cu–Mo–O alloy work-function plateaus at ∼6.25 eV). The work-function profiles follow a similar trend to the oxidation-state profiles and suggest that these profiles are correlated. The d6eV values of metals are 1.1 nm for V, 0.88 nm for Cu, 0.85 nm for Mo, 0.68 nm for Ni and 0.28 nm for Au (Note: d6eV is assigned as the thickness at which the work-function reaches 6.0 eV). The fact that work function depends on MoO3 thickness exemplifies the importance of oxide thickness and substrate choice for the energy-level alignment of molecular adsorbates.103
The results of Greiner et al.103 indicate that the interfacial MoO3 reduction occurs close to the metal/MoO3 interface function under two possible mechanisms: (i) charge transfer from the metal Fermi level to the MoO3 conduction band, and (ii) reduction of MoO3 driven by oxidation of the metal substrate.103 These two mechanisms assist comprehension concerning (i) how a noble metal, such as Au, can reduce Mo6+ within the first few nanometers of the Au/MoO3 interface, and (ii) how a reactive metal, such as V, can cause severe reduction of Mo6+, several nanometers from the interface of V/MoO3.
The results of Greiner et al.,103 demonstrate that the metal/metal-oxide interfaces critically influence the work function and band structure. It is indicated that the reduction of Mo6+, in interfacial metal/MoO3, results in changes to the MoO3 valence band structure so that the oxide semi-metal is close to the metal interface. The valence band changes can be illustrated in terms of electrons filling the previously empty Mo 4d states of MoO3; thereby resulting in donor states close to the Fermi level.103 The reduced Mo6+, to Mo5+ and Mo4+, also changes the work function of MoO3 to one close to the metal/MoO3 interface.103 The design and selection the appropriate oxide buffer thickness can be optimized from the above investigated data that is relevant for organic electronic devices.
Another study by Losego et al.110 synthesized an ultralow thermal conductivity nanolaminate system via a simple self-assembly method. They used organoclays produced by alkylammonium cation (R–NH3+) exchange with colloidal dispersed montmorillonite [(Na, Ca)0.33(Al, Mg)2(Si4O10)(OH)2·nH2O] clay sheets, pursued by solvent casting. The interface was investigated concerning the influence on the thermal conductivity of organoclay nanolaminates.
Fig. 25a shows a schematic of the sample geometry used to measure thermal conductivity of organoclay nanolaminates via time-domain thermoreflectance (TDTR). The results demonstrate that thermal conductivity is comparatively independent of nanolaminate spacing.110 A series resistance model was developed to better understand the behavior and influence of the interfacial thermal conductance for the organoclay nanolaminates:110 i.e.,
Fig. 25 (a) Schematic of the sample geometry used for TDTR measurements, (b) measured values for the cross-planar thermal conductivity of unmodified clay and organoclay nanolaminates prepared from varying alkylammonium chain lengths. The blue dashed line is an “effective medium” model that excepts interfacial conductance, the red dotted line supposes only interfacial conductance with G = 110 MW m−2 K−1, and the red solid line is a series resistance model that includes organic layer conductance (aluminosilicate conductance) and interfacial conductance with G = 150 ± 10 MW m−2 K−1.110 The thermal conductivity of WSe2 is plotted at its equivalent interface density (for reference).110 Inset images indicate the presumptions of 1 interface per unit cell for the unmodified clay and 2 interfaces per unit cell for the organoclays. The unmodified clay is plotted using its measured d spacing but has an assumed interface density of 0.85 interfaces per nm. Adapted with permission from data published in ref. 110 Copyright 2013, American Chemical Society. |
An interfacial thermal conductance of ∼150 MW m−2 K−1 was obtained for the organic/clay interface by Losego's model, which exhibits similar organic–inorganic interfaces. This interfacial conductance value presents opportunity for designing nanolaminate materials with lower thermal conductivities.138 The average thermal conductivity has been documented as 0.09 W m−1 K−1 for organoclay nanolaminates (Fig. 25b).110 This value is close to the lowest thermal conductivity ever reported for a dense solid; i.e., 0.06 W m−1 K−1 at room temperature for WSe2 films that were fabricated using more complex molecular beam evaporation techniques.139
The series resistance modeling, suggested by Losego et al., illustrate that the interface densities achieved in the organoclays (1–1.5 nm) are adequate to gain the ultralow thermal conductance. A material is thought to reach its lowest thermal conductivity when amorphous, since phonons are forced to transit by means of a random walk through atomic vibrations.140 Nevertheless, the amorphous limit can be surpassed once nanoscale features have been introduced.141–144 Interfaces in nanolaminates scatter phonons and provide additional thermal resistance to a system. Interface densities in nanolaminates can be sufficiently high to decrease the thermal conductivity to below the amorphous limit.145–147
Interfaces hinder phonon transport of thermal energy and its nanostructure can transform fully dense solids into ultralow thermal conductivity materials.110 Consequently, the above results suggest that organoclays can be developed as an experimental platform to explore nanolaminate materials and the role of interfaces in thermal physics.
Most 2D materials are usually produced by exfoliation or delamination of van der Waals solids,152,153 or by direct reaction between different elements or compounds.154–157 Recently, a selective extraction method has been designed to produce 2D materials derived from MAX phases.13 This method holds great interest because it operates at room temperature and has the capability of obtaining well-controlled morphologies and structures.13,148,158,159 As well, new materials can be produced by the selective extraction method that cannot be fabricated by other methods.13,148,158,159 For instance, MXenes and carbide-derived carbons (CDCs) are the most studied materials that are achieved by the selective extraction technique.160–162 Structures of type “A”, “M and A” have been selectively extracted from the MAX phases, MXene and CDCs materials, respectively.
These new extracted materials present a combined metal conductivity and hydrophilicity that could be used in energy storage (Li-ion batteries, supercapacitors) or in polymer nanocomposites.159–162 Recently, the selective extraction method, by electrochemical techniques at room temperature has been employed to extract the “M” element from Ti2SC MAX phase and produce an “AX” nanolaminated structures.163 The extraction method is applied at low potentials (0.3, 0.6, 1.0, 1.4 and 2.0 V) and ammonium fluoride (NH4F) aqueous solution was selected as the electrolyte. The bonding force between Ti and S in Ti2SC is weak, being van der Waals in nature, and S displays relatively little electrochemical interaction. Therefore, it is reasonable to predict that Ti would be easier to extract than S. The Ti extraction from Ti2SC results in carbon/sulfur (C/S) nanolaminates (Fig. 26a).
Fig. 26 (a) A schematic indicating the selective extraction of Ti (M) from Ti2SC (MAX phase) using electrochemical etching. (b) Raman spectra of Ti2SC before (bottom pattern) and after etching at various potentials. (c) Dependence of etched C/S layer thickness (after 1 hour) on etching voltage. (d) High-resolution TEM images of amorphous (left image) and graphene-like (right image) C/S flakes. Reproduced with permission from data published in ref. 163 Copyright 2015, John Wiley and Sons. |
The Raman spectra of Ti2SC after etching at 0.3, 0.6 and 1.0 V are almost the same and all of the Ti2SC peaks disappeared with just the C/S composite peaks revealed (Fig. 26b). The Raman spectra peaks are identified as C and small traces of TiO2 at the higher potential of 1.4 V. However, the intensity of TiO2 is enhanced at potential of 2.0 V, in addition to the C intensity. The TiO2 fabricated at higher voltages is thought to arise from the hydrolysis of the extracted Ti ions that are trapped within the C/S framework.163 It could also be suggested that both Ti and S are extracted from Ti2SC at the higher voltages. The potential rate also influenced the thickness of the C/S nanolaminate layers.
The results exhibited that after 1 h etching at 0.3, 0.6 and 0.8 V the thickness of C/S layers increased from 7 to 25 μm (Fig. 26c).163 HRTEM characterization shows that the fabricated C/S nanolaminates are significantly amorphous; however, 5% of the C/S displays a graphene-like structure (Fig. 26d).163 Li/S cells were fabricated and discharge capacities were measured for both materials to compare the fabricated C/S nanolaminates against sandwich-type graphene/S nanocomposites. The results illustrated that the fabricated C/S nanolaminates presented better cycling stability when tested under the same conditions. In additional to Ti2SC, some other MAX phases, such as Ti3AlC2, Ti3SnC2 and Ti2GeC were examined by the electrochemical etching process to produce “AX” nanolaminated structures.163 The products all revealed the “AX” layered structures, after extraction of the “M” constituent by using electrochemical etching. The established structure and characterization of nanolaminates derived from MAX phases reveal a research area with the potential of substantial important outcomes.
There are different binary oxides (HfO2, ZrO2, and TiO2, etc.) and ternary oxides (HfxAl1−xOy, and TixAl1−xOy, etc.) that are used in dielectric materials.164,175 However, these materials have a moderate dielectric constant (ε = 10 to 50), so that the desirable combination of high ε and low tanδ values is difficult for these single-phase materials.164 The reported ε values for perovskite oxides (BaTiO3, SrTiO3, etc.) are in ∼104 higher;175 however, perovskites have a complex structure that is difficult to grow onto Si substrates.175 Therefore, the nanolaminated forms of the binary oxides are an effective way to increase ε and reduce dielectric losses.164 Hence, the research studies have focused on mixed-phase or nanolaminate materials such as TiO2/Ta2O5, TiO2/Nb2O5, Nb2O5/Ta2O5, HfSOx/ZrSOx.164,175–177 The high ε value in nanolaminates (i.e., large permittivity of 330 at the frequency about 105 Hz in HfO2/ZnO nanolaminate when ZnO sublayer thickness is 13.6 nm178) (while relative permittivity of thin film HfO2 with 50 nm thickness at 100 Hz is ∼26 and for bulk ZnO is ∼8.66), was pointed out to be owing to the Maxwell–Wagner (MW) polarization via charge accumulation at the interfaces of insulating and semiconducting sublayers where space charge polarization increases ε.175 It is also reported that MW polarization can increase the measured value of ε in chemically homogenous systems.179
A dielectric with high ε and low leakage current density has been produced by the synthesis of amorphous nanolaminates of TiO2 (as the semiconductor) and Al2O3 (as an insulator) with individual layer thickness as low as 0.2 nm.175 An enhancement of ε value in the order of 104 up to a frequency of 10 kHz, dielectric losses smaller than 1 and extremely low leakage currents were reported for this nanolaminate.180,181 The high band gap of Al2O3 (8.8 eV) may hinder the leakage of the mobile charges and the tunneling can be controlled by the ultra-thin insulating layers for nanolaminate materials that use MW polarization. The amorphous character of nanolaminates also facilitates integration into existing technologies. Thus, this material may also be a promising candidate for integrated passive devices.175
Despite the overall advantages, the large-scale application of Li-ion batteries are limited by technological barriers, such as high cost, poor cycling stability and low electrical conductivity.188 A solution to these challenges employs 2D nanosheets and their composites with high conductive carbon materials for developing rate ability and cycling performance of Li-ion batteries.
Xie et al.189,190 have demonstrated that 2D nanosheets with an active and large surface area exhibit rapid Li storage and good cycling life. For instance, cobalt oxide (Co3O4) 2D nanosheets with a 1.5 nm thickness present a specific capacity equal to 812.8 mA h g−1 with little capacity loss per cycle and is an excellent anode material.189 Moreover, Li et al.191 reported that graphene/NiO 2D nanosheets could significantly increase the capacity, rate capability and cycling stability of Li-ion batteries. They have demonstrated that the enhancements in graphene/NiO are due to synergistic effects resulting from their good interfacial interaction that induces better performance rather than each individual component.191 Furthermore, Xie et al.192 illustrated that ultrathin graphene/β-Ni(OH)2 nanosheets could confer fast electron transfer and improve the high rate electrochemical performance. Against these attractive benefits of graphene/2D nanosheet composites, there are some drawbacks of graphene-based composites in Li-ion batteries.193
Two main limitations of graphene-based composites are: (i) preparation of graphene usually requires a time consuming and complex process through graphite exfoliation, or very high carbonization temperature is required from carbonaceous materials,194,195 and (ii) decoration of the graphene surface by active materials avoids their aggregation because of the lack of oxygen bridges between graphene and the active materials.191,195 Therefore, 2D nanosheets made from carbon-based composites could improve the Li-ion battery performance, compared to the complex synthesis process for graphene.
Xie et al.193 have synthesized sandwich-like nanosheets of carbon-anchored ultrathin TiO2 as a superior platform to achieve ultrafast Li storage kinetics and superior cycling stability. They employed a facile method and fabricated the ultrathin TiO2 nanosheets using a precursor of lamellar TiO2 and octylamine (C8H19N) (Fig. 27a). Firstly, lamellar hybrid nanosheets of TiO2–octylamine were fabricated by mixing precursors and hybridization at 180 °C.193 The lamellar hybrid nanosheets were then annealed at 450 °C for 2 h. The annealing process leads to growing anatase TiO2 and in situ carbonization of the intercalated organic components in the lamellar hybrid nanosheets (Fig. 27a).
Fig. 27 (a) A schematic from formation procedure of nanolaminate carbon-anchored ultrathin TiO2. (b) TEM image for the as-synthesized precursor lamellar TiO2–octylamine hybrid nanosheets (left image), TEM, inset HRTEM images (middle and right) corresponding elemental mapping for the nanolaminate carbon-anchored ultrathin TiO2. (c) A schematic illustration of advantage of using nanolaminate carbon-anchored ultrathin TiO2 nanosheets as the lithium ion battery electrode that facilitates fast lithium insertion/extraction, while anchored carbon provides conducting paths through the Ti–O–C bridge. Reproduced with permission from data published in ref. 193 Copyright 2014, The Royal Society of Chemistry (RSC). |
Fig. 27b exhibits TEM, HRTEM and elemental mapping of the as-synthesized precursor lamellar hybrid of TiO2–octylamine and sandwich-like nanosheets of carbon-anchored ultrathin TiO2. These images show that the as-synthesized precursor and sandwich-like morphologies have sizes of about 100 nm and 3–5 nm, respectively. The sandwich-like morphology also confirms that their structure did not change compared with the lamellar hybrid nanosheets of TiO2–octylamine. The corresponding elemental mapping analysis (Fig. 27b) also confirms the dispersion of the carbon component on the surface of the ultrathin TiO2 nanosheets. The cyclic voltammograms, cycling performance, efficacy of rate capacity and Nyquist values of carbon-anchored ultrathin TiO2 and bare TiO2 nanosheets have been studied and analyzed. The carbon-anchored ultrathin TiO2 nanosheets delivered a capacity of 101.9 mA h g−1 at discharge rates as high as 40 Coulombs at 6.8 A g−1; while the capacity after 1200 cycles at a large current density of 5 C is 150.4 mA h g−1.193 These are superior results compared to those reported for TiO2 nanosheets, with a reversible capacity of 82.2 mA h g−1 at the 630th cycle at a current density of 2000 mA g−1.196
Xie et al. reported that the ultrafast Li storage kinetics and higher capacity in carbon-anchored ultrathin TiO2 nanosheets may be attributed to a synergistic effect that improves Li storage.193 The TiO2 nanosheets with ultrathin thickness and large surface area assist to shorten the Li-ion diffusion length and increase the surface area contact with the electrolyte; hence, leading to faster kinetics and higher capacity193 (Fig. 27c). In addition, carbon derived from the carbonization of octylamine also increases conductivity and better provides a strain match. Thus, the carbon phase provides superior cycling stability and rate capability.193
Thin films are promising candidates for moisture and oxygen barriers. Meyer et al.202,203 employed the ALD method and fabricated Al2O3/ZrO2 nanolaminates to examine them as gas-diffusion barriers. The Al2O3/ZrO2 nanolaminates were grown at 80 °C. The measured permeation rate of the fabricated nanolaminate with 130 nm thickness, under a controlled environment of 70% humidity and 70 °C, was 4.7 × 10−5 g per m2 per day for water and 1.6 × 10−2 cm3 per m2 per day for oxygen.203 The Al2O3/ZrO2 nanolaminate with a ∼40 nm thickness also exhibited an ultralow WVTR of 3.2 × 10−4 g per m2 per day at 80% humidity and 80 °C.202 Choi et al.200 fabricated a SiO2/Al2O3 nanolaminate on a plastic substrate by RF-magnetron sputtering and then characterized its gas-diffusion barrier properties.200 The SiO2/Al2O3 gas barrier film of 480 nm thickness displayed a WVTR of 3.79 × 10−5 g per m2 per day in an environment of 20 °C and relative humidity of 50%. Lee et al. also demonstrated that the WVTR of the Al2O3 and ZrO2 layers were 9.5 × 10−3 g per m2 per day and 1.6 × 10−2 g per m2 per day, respectively, however when deposited alternatively with 1 cycle of each layer, the WVTR decreased to 9.9 × 10−4 g per m2 per day.204 The mechanism for the enhanced barrier efficiency can be attributed to the microstructure. The surface and cross-sectional characterization indicates suppressed formation of voids and extended crystals as a result of the alternating multilayer structure.203 In addition, the second phases of ZrO2 and SiO2 in the fabricated nanolaminate efficiently hamper the corrosion of Al2O3 by water as defined in the below chemical reactions.205
Al2O3(s) + 6H+(aq) + 3H2O(l) → 2[Al(H2O)3]3+(aq) |
Al2O3(s) + 2OH−(aq) + 3H2O(l) → 2[Al(OH)4]−(aq) |
Thus, the nanolaminated structures provide a mechanism for thin-film encapsulation of organic electronic devices and other applications because of the low permeation rates. In addition, the films exhibit close conformity to the intended device, which improves their tolerances and overall performance since encapsulation can be assured.
Therefore, both phonons and electrons contribute to the thermal conductivity. Hence, decoupling of these contributors for the thermal conductivity provides a design opportunity to suppress the phonon contribution (κ1) without a noticeable change in S and σ. This effect could be achieved by engaging phonon-blocking/electron-transmitting material structures.206 Hence, nanolaminated structures could boost the performance of thermoelectric materials because thermal conductivity may be suppressed without hindering the electrical transport properties.207,208 Ultrathin multilayer materials including organic/inorganic superlattices are attracting attention as potential thermoelectric materials for flexible power generation.206–208 These multilayered structures could block phonon transport by interfacial coupling and scattering effects without adversely influencing electric properties.207,208
Progress has been made on n-type organic/inorganic superlattice materials with flexible structures by using organic intercalation of layered transition metal dichalcogenides, TiS2.207 Koumoto et al. employed an electrochemical intercalation process and fabricated a TiS2/[(hexylammonium)x(H2O)y(DMSO)z] hybrid superlattice (Fig. 28a).207 Fig. 28b, indicates that the layers in the prepared nanolaminate hybrid superlattice are distorted and formed in a wavy structure. This wavy structure forms because the inorganic nanolayers tend to roll up and form nanotubes under a certain processing treatments.207 The equilibrium molecular dynamics simulations have indicated that the thermal conductivity is less than the experimental observation.207 The lower simulated thermal conductivity may be attributed to the wavy structure layers which provide an additional mechanism for reducing thermal conductivity of the organic/inorganic multilayer structure.
Fig. 28 (a) A schematic indicating synthesis of TiS2-based inorganic/organic superlattices. Firstly, TiS2 single crystal electrochemically intercalated into a TiS2[(HA)x(DMSO)y] superlattice where a bilayer structure of the hexylammonium ions was formed due to DMSO stabilization. After that, superlattice of TiS2[(HA)x(H2O)y(DMSO)z] formed by the solvent exchange process after immersion in water where the hexylammonium ions change to a monolayer configuration. (b) HAADF-STEM images of the TiS2[(HA)x(H2O)y(DMSO)z] hybrid superlattice displaying a wavy structure. (c–f) Electrical conductivity, thermal conductivity, Seebeck coefficient and ZT values of TiS2 single crystal and TiS2[(HA)0.08(H2O)0.22(DMSO)0.03], respectively. Reproduced with permission from data published in ref. 207 Copyright 2015, Macmillan Publishers Ltd: Nature Materials. |
The electrical conductivity, thermal conductivity, Seebeck coefficient and ZT values of TiS2 single crystal and TiS2[(HA)0.08(H2O)0.22(DMSO)0.03] are shown in Fig. 28c–f. An electrical conductivity of 790 S cm−1 and a power factor of 0.45 mW m−1 K−2 with an in-plane lattice thermal conductivity of 0.12 ± 0.03 W m−1 K−1 are obtained in the multilayer hybrid superlattice. The ZT value at 373 K for this material is 0.28, which is the highest quantity for n-type flexible composites. It is a challenge to obtain high values of ZT for n-type doping organic semiconductors due to their low electron affinity.207 The value of ZT in the multilayer n-type structure is close to the most promising PEDOT-PSS p-type organic material that exhibits the highest ZT of 0.42.208
Many other nanolaminate structures such as Bi2Te3, NbS2, TaS2, VS2, CrS2, MoS2 and NaxCoO2 could also be intercalated to make both p- and n-type thermoelectric materials.207 The ultimate selections of organic molecules and inorganic materials with different sizes and functional groups could provide strategies to optimize the thermoelectric efficiency.207
Ti3SiC2 can be added onto ceramic armor210 to act (i) as a laminar material combined with a metal layer or (ii) as a functionally gradient material (FGM) with other ceramics. Such materials technology may replace currently used ceramic working elements by conferring a benefit of improved crack resistance and fracture toughness. For instance, Ti3SiC2–Al2O3 nanolaminates demonstrate high damage energy and increased fracture toughness.210 The design of a FGM requires all of the constituents to be well defined in terms of elastic properties (i.e., Young's and shear module and Poisson's ratio) and thermal properties (i.e., thermal expansion coefficient and thermal conductivity).
Nanolaminates may be coated onto, or manufactured into, thin-shell curved mirrors for high resolution imaging in visible and infrared light for terrestrial or extra-terrestrial applications to create lightweight adaptive imaging mirrors.211 These mirrors may have diameters of the order of a meter and include metal film reflectors on nanolaminate substrates supported by multiple distributed piezoceramic “piston”-type actuators for micron-level control.211 Whilst the conventional glass mirrors with equal size and precision have mass densities between 50 and 150 kg m−2, the nanolaminate mirrors, including not only the reflector/shell portions but also the actuators and the backing structures needed to react the actuation forces, would have mass densities that may approach ∼5 kg m−2.211 Moreover, whereas fabrication of a traditional glass mirror of equivalent precision requires several years, the reflector/shell portion of a nanolaminate mirror could be fabricated in less than a week, and its actuator system could be fabricated in 1 to 2 months.211
MXene materials are important 2D structures that can be included in the ceramic- and magnetic-based category of nanolaminates. They are 2D transition metal carbides and/or nitrides that are fabricated by extracting “A” layer from MAX phases and which integrate metal and ceramic properties.13 MXenes by exhibiting good electronic conductivity and a 2D morphology are suitable for storage energy devices, such as Li-ion batteries, hybrid cells and supercapacitors. The experimental development of MXene materials can address current issues13 such as (i) a full understanding of surface chemistry, (ii) characterization of chemical and thermal stabilities under different environmental conditions, (iii) electronic, magnetic, optical, thermal and mechanical characterization of single layers of the MXenes, (iv) exploring potential applications (i.e., electronic, magnetic, sensors, etc.), and (v) investigation of new MXenes.
In addition to MXene phases, applied and fundamental research on other ceramic-based nanolaminates for energy storage (dielectrics of capacitors,164 batteries193,212), gas-diffusion barriers200 and energy conversion (thermoelectric)206,207 is underway. For instance, sandwich-like carbon-anchored ultrathin TiO2 nanosheets, proposed as a superior platform to achieve ultrafast Li storage kinetics and superior cycling stability, has been explored.193 Progress has also been reported on n-type organic/inorganic superlattice materials with flexible structures that have been formed by organic intercalation of layered transition metal dichalcogenides TiS2, which will be of interest for thermoelectric applicable.207
There are numerous applications in many fields of technology and engineering for nanolaminate composite materials. It is anticipated that current and future nanolaminates will require continued identification and exploration for specific applications. The expansion of nanolaminate structures face several challenges, such as (i) better understanding of structure–property relationships that predict the performance of a given nanolaminate structure, and (ii) identification of large-scale, low-cost and simple methods for industrial production. Overcoming these barriers will provide nanolaminates with excellent assure for applications in nanoelectronics, sensing, catalysis, gas separation and energy related areas.
This journal is © The Royal Society of Chemistry 2016 |