Exfoliation (chemistry)

Last updated

Exfoliation is a process that separates layered materials into nanomaterials by breaking the bonds between layers using mechanical, chemical, or thermal procedures.

Contents

While exfoliation has historical roots dating back centuries, significant advances and widespread research gained momentum after Novoselov and Geim's discovery of graphene using Scotch tape in 2004. Their Nobel Prize-winning research primarily relied on mechanical exfoliation for the production of graphene which sparked an immediate interest in the exfoliation process. Today, exfoliation is regarded as the most widely used nanomaterial production technique.

Exfoliation typically involves breaking weak bonds called van der Waals bonds to create two-dimensional materials, such as graphene or transition metal dichalcogenide monolayers. While various reversible chemical processes, such as intercalation can disrupt the weak bonds in a lamellar structure and introduce guest species, many of them fail to produce single-sheet materials as the processes are not strong enough to cancel the interlayer attractions. [1] However, during exfoliation, the high energy input leads to an extreme bond-breaking process that irreversibly separates the layers into single sheets. Lately, it has been shown that if the energy input is substantial enough, the procedure can even break much stronger, bonds such as metallic or ionic bonds to create non-van der Waals materials like hematene or other nanoplatelets. [2]

In recent years, exfoliation has found practical applications in a wide range of fields, from electronics to biomedical and beyond. It plays a vital role in creating advanced materials with properties tailored for specific uses, such as high-performance electronics, efficient energy storage devices, and lightweight yet robust materials for aerospace applications. This versatility and adaptability make exfoliation a crucial technique in cutting-edge material research and various industrial sectors.

History

While the use of exfoliation can be traced back to ancient Chinese and Mayan pottery, the earliest scientific work involving exfoliation was the production of vermiculite by Thomas H. Webb, in 1824. [3] [4] However, during this early period, no substantial research was conducted to understand the nature of the mechanisms that facilitated these reactions. Arguably, the first research that delved into the mechanism of the process rather than its usage was Brodie's work, which revealed that certain acids produced lamellar carbon structures in 1855. [5] Despite this discovery, extensive research on the topic did not immediately follow. The exfoliation concepts we have today were not developed until the realization that graphite absorbed alkali metals in 1926. [6] This discovery laid the groundwork for a more solid theoretical framework, enabling scientists to apply the method in their production processes. One method that made use of this theoretical background and eventually led to further development of the process as a production technique was Rüdorff and Hoffman's work, which introduced an electrochemical method for exfoliation in 1938. [7] The development of electrochemical exfoliation piqued the interest of more researchers and more people started regarding the process as a production technique. One of the most notable examples of the success of the method as a mass production technique was the invention of the first commercial lithium carbon fluoride batteries in 1973. [8]

The real turning point for exfoliation research came in 2004 when Novoselov and Geim isolated graphene through mechanical exfoliation using Scotch tape. This innovative research earned them the Nobel Prize in Physics in 2010, reigniting the interest in exfoliation methods. [9] In subsequent years, numerous processes were developed for more precise manufacturing and higher yields. While most of the exfoliation research focused on graphite and graphene during the last few decades, recently, the rather difficult processing of graphene and its lack of an obvious band structure led many research groups to begin working on different elements to utilize exfoliation for the production of other nanomaterials. [10]   One of the most significant breakthroughs of this new wave of research has been the discovery of non-van der Waals nanoplatelets. This discovery demonstrated that exfoliation could occur without relying on weak bonds, which opened up new and promising applications in the industry.

Types of Exfoliation

The exfoliation process is typically applied to lamellar structures with weak bonds. While these bonds are weak enough to be easily broken by an external force, they are strong enough to not separate into single layers. [11] In order to separate the material into single layers, the attraction between consecutive layers must be overcome. As the interest in exfoliated material research surged many researchers started to develop new and better ways to overcome these interlayer attractions. Despite the high number of methods it is possible to classify them into three distinct categories based on the source of energy used in the process: mechanical, chemical, and thermal exfoliation.

Mechanical Exfoliation

In mechanical exfoliation, external forces act upon the material, breaking the bonds due to the stress accumulated within the material. [12] Depending on the intensity and the specific nature of the phenomenon, these forces break the van der Waals forces, separating materials into 2D nanostructures. Sometimes, a solvent is introduced to the material to facilitate complete breakdown, as liquid environments significantly reduce bond strength compared to vacuum conditions. [9]

While mechanical exfoliation is effective in separating the layers, it lacks predictability and systematic results. The process requires repetition until individual layers are achieved. To obtain consistent nanomaterials with specific properties, experimentation and fine-tuning of conditions based on the results is required. [12] Therefore, mechanical exfoliation techniques are rather empirical, and most mathematical models rely on empirical results rather than the ab-initio calculations. [13]

The original method proposed by Novoselov and Geim, micromechanical cleavage,  was essentially a mechanical exfoliation method. Consequently, mechanical exfoliation methods were developed more rapidly than the others. Major mechanical exfoliation methods include micromechanical cleavage and liquid phase separation.

Micromechanical Cleavage

Micromechanical Cleavage is the original graphene production method proposed by Novoselov and Geim. This process involves using sticky tape to get graphite samples and separating the layers until getting a single layer. [9] Although the process yielded high-purity single-layered materials, it fell out of favor quickly as it required several repetitions for a single layer of graphene and was likely to damage the graphene layers during the process. [13]

Liquid Phase Separation

Liquid phase separation is one of the most widely used exfoliation methods. Its high yields, high purity, and scalability make it one of the most preferred exfoliation methods. It works by providing a liquid medium for the mechanical exfoliation methods. The liquid medium significantly reduces the strength of the bonds in the material compared to vacuum conditions, making it easier for mechanical forces to break the weak bonds in the material. [9] However, due to the interfacial tension forces, liquid phase separation does not always yield uniform results. When the tension forces do not balance out, graphene single layers may break due to the tension forces. To achieve relatively uniform results, the overall energy of the system must be minimized. The best way to optimize this condition is to use solvents with similar surface tensions to the material of interest. [14] Liquid phase separation utilizes various external forces to break the van der Waals forces. The most widely used liquid phase separation methods include sonication, which uses sound waves, and shearing, which uses shear forces.

Sonication

The sonication method utilizes ultrasonic sound waves to create micrometer-sized bubbles in liquid environments. When these bubbles reach a critical size, they collapse with an instantaneous temperature of 5000 K, a local pressure of 20MPa, and a heating/cooling rate up to 109 K s−1.These sudden physical differences create shock waves that can act on lamellar materials and break the weak forces in between the layers. [15] [16] Although sonication is a long-known laboratory technique, its implementation into graphene exfoliation was in 2008 and it led to liquid exfoliation becoming the predominant technique. [16]

While sonication is generally used as an exfoliation method on its own, it is also used as a further processing method to perfect the nanoflakes created with other exfoliation methods. Therefore it is a common technique used in combination with the other methods. One disadvantage of sonication is the reaction time though. A complete exfoliation reaction may take days to finish. However, prolonged exfoliation times allow the creation of more stable solutions, making long sonication times favorable for obtaining purer, defect-free products. Nanomaterials created with sonication yield 1.5 times larger unperturbed particle size. [15]

Shearing

The shearing method makes use of lab mixers to exfoliate lamellar structures into single-layered nanomaterials. Lab mixers create a sufficient shear force that allows consecutive layers of the material to slip over each other. [17] Which produces massive quantities of highly pure material.  Although the shearing method was widely used as a further processing method to break up relatively larger clusters of nanomaterials into single layers, before, in 2010, it was introduced as a direct method to exfoliate graphite into graphene. [18] Later studies confirmed the applicability of the method to other lamellar materials such as h-BN, MoS2, WS2, MoSe2, and MoTe2.

While this method has a high yield and purity among the other exfoliation methods, its known linear relations with concentration, mixing time, rotor speed, rotor diameter, and inverse relation with liquid volume, gives one of the best controllability out of all the exfoliation methods. [17] This innovative procedure has been adapted for household kitchen mixers, significantly reducing the costs and complexity of the exfoliation methods, thereby sparking another wave of research in layered structures. [19] [20]

Chemical Exfoliation

Chemical exfoliation employs the intercalation process to separate material layers. During intercalation, guest ions or free electrons are introduced to the layers, disrupting the bond structure and forming new bonds. [1] For example, in the case of van der Waals forces, which are common in chemical exfoliation, positive and negative regions are induced, attracting ions. Given that the bonds between layers are weak, they tend to break, forming new, stronger bonds with these ions. [9] Typically, these stronger bonds lead to the creation of functional groups that significantly reduce interlayer attractions. At this stage, the interlayer attraction becomes low, and thanks to the ability of the functional groups to decompose with further processing, the layers can be easily separated. [9]

Chemical exfoliation's scalability advantages over other production methods have made it one of the most widely used techniques. In addition to its scalability, the variety of chemicals available plays an important role in encouraging researchers to explore various production methods. Chemical exfoliation is also commonly used in combination with mechanical and thermal exfoliation methods to obtain purer results. The most widely used chemical exfoliation methods are chemical vapor deposition, graphite oxide reduction, and electrochemical exfoliation.

Chemical Vapor Deposition

First introduced in 2008, chemical vapor deposition emerged as one of the most popular methods for graphene exfoliation. This method utilizes a transition metal film as a base layer and exposes it to hydrocarbons at high temperatures(900-1000°C) and ambient pressure. During the process, hydrocarbon decomposes, and carbon atoms form one to ten layers of graphene flakes over the metal film. The metal film is then cooled down at a determined rate to achieve specific particle sizes. This process is especially useful for applications such as circuit drawing and surface-based applications of graphene, including the production of photovoltaic cells. [21] [22] Although the method was widely used until the last decade, its relatively expensive process has been replaced by other methods. However, there is still ongoing research to further develop the process for more efficient use with various materials. [22]

Oxide Reduction

The oxide reduction method is particularly widely used with graphite to create graphene. It involves introducing oxide functional groups into the lamellar structure, which doubles the distance between graphite layers and reduces van der Waals attractions. [9] These functional groups are then removed using reductants, resulting in single graphene layers from the graphite, which can now be easily exfoliated due to reduced van der Waals attractions. This method is especially valuable for fine-tuning the band gap properties of graphene, which naturally lacks a band gap. [23] [9]

While this method was widely used over the last decade, its impurity levels led to its decline in popularity. The presence of a large number of holes and defects made the produced graphene unsuitable for electronics, and the chemicals used were hazardous. [15] In 2014, a research group succeeded in isolating graphene layers without the use of oxidants, significantly increasing the purity of the samples and eliminating the need for further processing of the products. [5] This advancement is expected to reignite interest in oxide reduction exfoliation.

Electrochemical Exfoliation

One of the most promising exfoliation methods is electrochemical exfoliation, which has been popular among researchers since its introduction in 2008. [15] This method is mainly based on 20th-century studies on electrolysis and electrochemical intercalation. [24] Electrochemical exfoliation makes use of potential differences between a lamellar structured electrode and a platinum electrode to attract oppositely charged ions to the electrodes. These accumulations trigger the intercalation process in the material and ultimately result in the complete exfoliation of the material into single nanomaterial layers. [9] However, intercalation is not always the only reaction mechanism, as sometimes bubbles are observed depending on the solvent and electrolyte used. These bubbles also facilitate exfoliation by creating a similar effect to the sonication method. [17]

The process might be called cathodic or anodic exfoliation, depending on which electrode is the lamellar structured electrode. Cathodic exfoliation requires an organic solvent medium with a lithium or alkylammonium electrolyte, while anodic exfoliation can be done with water and strong electrolytes. Anodic exfoliation is more efficient than cathodic exfoliation, as it forms oxide and hydroxide functional groups, significantly increasing intercalation in the material. However, anodic exfoliation also results in impure products, so the choice between the two methods depends on the specific application. [15] [12] Electrochemical exfoliation products may also require further processing.

Unlike liquid exfoliation, electrochemical exfoliation eliminates most of the chemical reactions involved, resulting in purer products. This method increases scalability, controllability, and decreases contamination and reaction time for the exfoliated material. Therefore, many researchers aim to implement the method into the industry for the mass production of carbon nanomaterials and transition metal dichalcogenide monolayers. [25]

Thermal Exfoliation

Thermal exfoliation uses heat as a source of energy for the exfoliation process. Despite heat being such a fundamental energy for most of the other chemical processes its use in exfoliation is relatively recent. Most thermal exfoliation methods have the same approach; chemically intercalated lamellar structures are subjected to extreme temperatures to decompose the functional groups created through chemical methods. The decomposition of these functional groups generates gases that build up pressure between layers, countering the van der Waals attractions between material layers. [9] When well-chosen functional group/temperature combinations are used, complete separation of the layers occurs.

One advantage of thermal exfoliation methods over others is their higher production rate, a crucial property for mass production applications. [21] Additionally, their reaction times are the shortest among all exfoliation methods. A process that might take days to complete with mechanical exfoliation can be finished within seconds using thermal exfoliation methods. [9] However, reduced reaction time and higher yields come at the cost of reduced control over particle size due to the nature of the process. Therefore, the process still lacks the optimization and reproducibility required by the industry. [14] Today, the most widely used thermal methods are high-temperature, low-temperature, and microwave exfoliation methods.

High Temperature Thermal Exfoliation

High-temperature thermal exfoliation employs temperatures above 550°C to decompose functional groups. The biggest advantage of this method is its short reaction times. An exfoliation process that might take days to complete with mechanical exfoliation can be done in a matter of seconds through high-temperature thermal exfoliation. However, decreased reaction times come at the price of impure products. Due to the extremely high temperatures, operation costs increase significantly. Moreover, the carbon dioxide and water vapors produced during the decomposition of oxide groups react with the material, causing defects and impurities in the material. [26]

Low Temperature Thermal Exfoliation

Low-temperature thermal exfoliation aims to retain the benefits of high-temperature thermal exfoliation while avoiding unexpected outcomes such as high costs and impurities. For this purpose, low-temperature thermal exfoliation employs relatively lower temperatures of 200°C-550°C to decompose the functional groups. [9] These temperatures yield purer results than high-temperature thermal exfoliation because the chemicals produced at this temperature do not readily react with the layered material itself. [26] Although this decrease in temperature affects reaction times, it is usually favored to achieve purer results. Even though the reaction time is shorter in low-temperature thermal exfoliation compared to high-temperature, it is still significantly shorter than in other methods. Additionally, low-temperature thermal exfoliation allows for fine-tuning of the bandgap properties of materials, making it an ideal method for electronic applications. [23]

Microwave Irradiation Exfoliation:

Microwave Irradiation Exfoliation is another exfoliation method that would decrease the complexity of the exfoliation experiments greatly. First utilized for the production of exfoliated graphite, it was later adapted for other nanomaterials. In the microwave irradiation exfoliation method, materials partially intercalated through chemical processes are exposed to microwave radiation. Ions and molecules trapped between layers absorb microwaves, leading to local temperature changes. These local changes trigger significant physical and chemical phenomena that result in complete exfoliation of the lamellar material. [9] [17] Due to reduced costs and high efficiency, microwave irradiation exfoliation is one of the most popular exfoliation methods. The method also provides higher yields with pure results within shorter timeframes. [17] Although microwave irradiation exfoliation has great benefits there is still some ambiguity in the mechanisms of this method as the products of the method are reported to be able to get exfoliated again through chemical exfoliation. [9]

Applications

Ever since the isolation of graphene, exfoliation has been the most common and reliable method for creating graphene, with the ongoing development of new techniques to optimize the process. As graphene finds increasing applications in various areas of electronics, the quest for an optimized industrial production method for graphene becomes more significant. Currently, graphene is projected to play a crucial role in the production of low-cost solar cells, energy storage systems, and sensors. Therefore, various forms of graphene, from liquid suspensions to dispersions, coatings to dust, are necessary for implementation in industrial production methods. [13] In addition to the graphene, the exfoliation process enables the production of various other carbon allotropes, with the most important ones being carbon nanotubes and carbon quantum dots. These materials are also expected to create billion-dollar industries, and as a result, commercialization of these materials are anticipated to show advancements in exfoliation methods.

Although graphene is expected to be one of the most important materials in the future, there are still some disputes about some of its applications. The challenging processing of graphene and its lack of an obvious band structure have led many researchers to explore new uses of the exfoliation methods. This shift has recently increased research into efficient production methods for transition metal dichalcogenide (TMD) monolayers significantly. [10] TMD monolayers have band gaps ranging from insulators to semiconductors, thanks to their quantum confinement effects. Therefore, they are expected to have significant applications in the near future, particularly with the further development of optoelectronics. [13] Currently, TMD monolayers find applications in electronic devices such as solar cells, photodetectors, light-emitting diodes, and phototransistors. There is also a growing interest in their use in power storage systems, such as batteries and supercapacitors. [27] Since exfoliation is TMD monolayers' most common production technique, it is projected that TMD monolayers' potential commercialization will require extensive use of exfoliation methods, eventually creating new applications for exfoliation.

Theoretically, exfoliation requires the presence of weak bonds. However, recent studies have shown that even materials with metallic and ionic bonds can be exfoliated with the proper procedures. The materials created through these methods are called non-van der Waals nanoplatelets. One notable non-van der Waals material is the Hematane which is a single sheet of hematite, the most abundant form of iron ore. [2] Hematane is known to have interesting photocatalytic properties due to its modified bandgap properties, offering potential applications in energy storage, optoelectronics, and biomedicine. [28] Since one of the most common ways to create hematite is through liquid phase separation, applications of hematite would increase the interest in exfoliation.

Related Research Articles

<span class="mw-page-title-main">Boron nitride</span> Refractory compound of boron and nitrogen with formula BN

Boron nitride is a thermally and chemically resistant refractory compound of boron and nitrogen with the chemical formula BN. It exists in various crystalline forms that are isoelectronic to a similarly structured carbon lattice. The hexagonal form corresponding to graphite is the most stable and soft among BN polymorphs, and is therefore used as a lubricant and an additive to cosmetic products. The cubic variety analogous to diamond is called c-BN; it is softer than diamond, but its thermal and chemical stability is superior. The rare wurtzite BN modification is similar to lonsdaleite but slightly softer than the cubic form.

<span class="mw-page-title-main">Graphite</span> Allotrope of carbon, mineral, substance

Graphite is a crystalline form of the element carbon. It consists of stacked layers of graphene. Graphite occurs naturally and is the most stable form of carbon under standard conditions. Synthetic and natural graphite are consumed on a large scale for uses in pencils, lubricants, and electrodes. Under high pressures and temperatures it converts to diamond. It is a good conductor of both heat and electricity.

<span class="mw-page-title-main">Molybdenum disulfide</span> Chemical compound

Molybdenum disulfide is an inorganic compound composed of molybdenum and sulfur. Its chemical formula is MoS
2
.

<span class="mw-page-title-main">Graphene</span> Hexagonal lattice made of carbon atoms

Graphene is an allotrope of carbon consisting of a single layer of atoms arranged in a hexagonal lattice nanostructure. The name is derived from "graphite" and the suffix -ene, reflecting the fact that the graphite allotrope of carbon contains numerous double bonds.

<span class="mw-page-title-main">Intercalation (chemistry)</span> Reversible insertion of an ion into a material with layered structure

In chemistry, intercalation is the reversible inclusion or insertion of a molecule into layered materials with layered structures. Examples are found in graphite and transition metal dichalcogenides.

<span class="mw-page-title-main">Bismuth telluride</span> Chemical compound

Bismuth telluride is a gray powder that is a compound of bismuth and tellurium also known as bismuth(III) telluride. It is a semiconductor, which, when alloyed with antimony or selenium, is an efficient thermoelectric material for refrigeration or portable power generation. Bi2Te3 is a topological insulator, and thus exhibits thickness-dependent physical properties.

<span class="mw-page-title-main">Nanobatteries</span> Type of battery

Nanobatteries are fabricated batteries employing technology at the nanoscale, particles that measure less than 100 nanometers or 10−7 meters. These batteries may be nano in size or may use nanotechnology in a macro scale battery. Nanoscale batteries can be combined to function as a macrobattery such as within a nanopore battery.

As the world's energy demand continues to grow, the development of more efficient and sustainable technologies for generating and storing energy is becoming increasingly important. According to Dr. Wade Adams from Rice University, energy will be the most pressing problem facing humanity in the next 50 years and nanotechnology has potential to solve this issue. Nanotechnology, a relatively new field of science and engineering, has shown promise to have a significant impact on the energy industry. Nanotechnology is defined as any technology that contains particles with one dimension under 100 nanometers in length. For scale, a single virus particle is about 100 nanometers wide.

Exfoliated graphite nano-platelets (xGnP) are new types of nanoparticles made from graphite. These nanoparticles consist of small stacks of graphene that are 1 to 15 nanometers thick, with diameters ranging from sub-micrometre to 100 micrometres. The X-ray diffractogram of this material would resemble that of graphite, in that the 002 peak would still appear at ~26o 2 theta. However, the peak would appear considerably smaller and broader. These features indicate that the interplanar distance in exfoliated graphite is similar to that of the parent graphite, but the stack size is small. Since xGnP is composed of the same material as carbon nanotubes, it shares many of the electrochemical characteristics, although not the tensile strength. The platelet shape, however, offers xGnP edges that are easier to modify chemically for enhanced dispersion in polymers.

<span class="mw-page-title-main">Graphite oxide</span> Compound of carbon, oxygen, and hydrogen

Graphite oxide (GO), formerly called graphitic oxide or graphitic acid, is a compound of carbon, oxygen, and hydrogen in variable ratios, obtained by treating graphite with strong oxidizers and acids for resolving of extra metals. The maximally oxidized bulk product is a yellow solid with C:O ratio between 2.1 and 2.9, that retains the layer structure of graphite but with a much larger and irregular spacing.

<span class="mw-page-title-main">Jonathan Coleman (physicist)</span> Irish physicist

Jonathan Coleman is the Erasmus Smith's Professor of Natural and Experimental Philosophy in the School of Physics and a Principal Investigator in CRANN at Trinity College Dublin. Coleman's research focuses on solution-processing of nanomaterials and their use in applications. He is most well known for the development of liquid phase exfoliation, a widely used method for preparing two-dimensional nanosheets.

<span class="mw-page-title-main">Transition metal dichalcogenide monolayers</span> Thin semiconductors

Transition-metal dichalcogenide (TMD or TMDC) monolayers are atomically thin semiconductors of the type MX2, with M a transition-metal atom (Mo, W, etc.) and X a chalcogen atom (S, Se, or Te). One layer of M atoms is sandwiched between two layers of X atoms. They are part of the large family of so-called 2D materials, named so to emphasize their extraordinary thinness. For example, a MoS2 monolayer is only 6.5 Å thick. The key feature of these materials is the interaction of large atoms in the 2D structure as compared with first-row transition-metal dichalcogenides, e.g., WTe2 exhibits anomalous giant magnetoresistance and superconductivity.

In materials science, the term single-layer materials or 2D materials refers to crystalline solids consisting of a single layer of atoms. These materials are promising for some applications but remain the focus of research. Single-layer materials derived from single elements generally carry the -ene suffix in their names, e.g. graphene. Single-layer materials that are compounds of two or more elements have -ane or -ide suffixes. 2D materials can generally be categorized as either 2D allotropes of various elements or as compounds.

A rapidly increasing list of graphene production techniques have been developed to enable graphene's use in commercial applications.

<span class="mw-page-title-main">Boron nitride nanosheet</span>

Boron nitride nanosheet is a crystalline form of the hexagonal boron nitride (h-BN), which has a thickness of one atom. Similar in geometry as well as physical and thermal properties to its carbon analog graphene, but has very different chemical and electronic properties – contrary to the black and highly conducting graphene, BN nanosheets are electrical insulators with a band gap of ~5.9 eV, and therefore appear white in color.

Graphene is the only form of carbon in which every atom is available for chemical reaction from two sides. Atoms at the edges of a graphene sheet have special chemical reactivity. Graphene has the highest ratio of edge atoms of any allotrope. Defects within a sheet increase its chemical reactivity. The onset temperature of reaction between the basal plane of single-layer graphene and oxygen gas is below 260 °C (530 K). Graphene combusts at 350 °C (620 K). Graphene is commonly modified with oxygen- and nitrogen-containing functional groups and analyzed by infrared spectroscopy and X-ray photoelectron spectroscopy. However, determination of structures of graphene with oxygen- and nitrogen- functional groups requires the structures to be well controlled.

<span class="mw-page-title-main">Discovery of graphene</span>

Single-layer graphene was first unambiguously produced and identified in 2004, by the group of Andre Geim and Konstantin Novoselov, though they credit Hanns-Peter Boehm and his co-workers for the experimental discovery of graphene in 1962; while it had been explored theoretically by P. R. Wallace in 1947. Boehm et al. introduced the term graphene in 1986.

A graphene helix, similar to the carbon nanotube, is a structure consisting of a two-dimensional sheet of graphene wrapped into a helix. These graphene sheets can have multiple layers, called multi-walled carbon structures, that add to these helices thus increasing their tensile strength but increasing the difficulty of manufacturing. Using van der Waals interactions it can make structures within one another.

<span class="mw-page-title-main">Layered materials</span>

In material science, layered materials are solids with highly anisotropic bonding, in which two-dimensional sheets are internally strongly bonded, but only weakly bonded to adjacent layers. Owing to their distinctive structures, layered materials are often suitable for intercalation reactions.

First demonstrated in 2008, liquid-phase exfoliation (LPE) is a solution-processing method which is used to convert layered crystals into 2-dimensional nanosheets in large quantities. It is currently one of the pillar methods for producing 2D nanosheets. According to IDTechEx, the family of exfoliation techniques which are directly or indirectly descended from LPE now make up over 60% of global graphene production capacity.

References

  1. 1 2 Whittingha, Stanley M. (2012-12-02). Intercalation Chemistry. Elsevier. ISBN   978-0-323-14040-9.
  2. 1 2 Puthirath Balan, Aravind; Radhakrishnan, Sruthi; Woellner, Cristiano F.; Sinha, Shyam K.; Deng, Liangzi; Reyes, Carlos de los; Rao, Banki Manmadha; Paulose, Maggie; Neupane, Ram; Apte, Amey; Kochat, Vidya; Vajtai, Robert; Harutyunyan, Avetik R.; Chu, Ching-Wu; Costin, Gelu (July 2018). "Exfoliation of a non-van der Waals material from iron ore hematite". Nature Nanotechnology. 13 (7): 602–609. doi:10.1038/s41565-018-0134-y. ISSN   1748-3395.
  3. "Clay-organic interactions". Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences. 311 (1517): 315–332. 1984-06-14. doi:10.1098/rsta.1984.0031. ISSN   0080-4614.
  4. Nicolosi, Valeria; Chhowalla, Manish; Kanatzidis, Mercouri G.; Strano, Michael S.; Coleman, Jonathan N. (2013-06-21). "Liquid Exfoliation of Layered Materials". Science. 340 (6139). doi:10.1126/science.1226419. hdl: 2262/69769 . ISSN   0036-8075.
  5. 1 2 Kovtyukhova, Nina I.; Wang, Yuanxi; Berkdemir, Ayse; Cruz-Silva, Rodolfo; Terrones, Mauricio; Crespi, Vincent H.; Mallouk, Thomas E. (November 2014). "Non-oxidative intercalation and exfoliation of graphite by Brønsted acids". Nature Chemistry. 6 (11): 957–963. doi:10.1038/nchem.2054. ISSN   1755-4349.
  6. McKinnon, W. R.; Haering, R. R. (1983), White, Ralph E.; Bockris, J. O’M.; Conway, B. E. (eds.), "Physical Mechanisms of Intercalation", Modern Aspects of Electrochemistry: No. 15, Boston, MA: Springer US, pp. 235–304, doi:10.1007/978-1-4615-7461-3_4, ISBN   978-1-4615-7461-3 , retrieved 2023-11-09
  7. Hofmann, Ulrich; Rüdorff, Walter (1938-01-01). "The formation of salts from graphite by strong acids". Transactions of the Faraday Society. 34: 1017–1021. doi:10.1039/TF9383401017. ISSN   0014-7672.
  8. Ikram, Muhammad; Raza, Ali; Ali, Sarfraz; Ali, Salamat (2020-12-04), "Electrochemical Exfoliation of 2D Advanced Carbon Derivatives", 21st Century Advanced Carbon Materials for Engineering Applications - A Comprehensive Handbook, IntechOpen, ISBN   978-1-78985-924-9 , retrieved 2023-11-09
  9. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Cai, Minzhen; Thorpe, Daniel; Adamson, Douglas H.; Schniepp, Hannes C. (2012-11-20). "Methods of graphite exfoliation". Journal of Materials Chemistry. 22 (48): 24992–25002. doi:10.1039/C2JM34517J. ISSN   1364-5501.
  10. 1 2 Zhang, Qingyong; Mei, Liang; Cao, Xiehong; Tang, Yuxin; Zeng, Zhiyuan (2020-08-11). "Intercalation and exfoliation chemistries of transition metal dichalcogenides". Journal of Materials Chemistry A. 8 (31): 15417–15444. doi:10.1039/D0TA03727C. ISSN   2050-7496.
  11. Gao, Enlai; Lin, Shao-Zhen; Qin, Zhao; Buehler, Markus J.; Feng, Xi-Qiao; Xu, Zhiping (2018-06-01). "Mechanical exfoliation of two-dimensional materials". Journal of the Mechanics and Physics of Solids. 115: 248–262. doi: 10.1016/j.jmps.2018.03.014 . ISSN   0022-5096.
  12. 1 2 3 Abdelkader, A. M.; Cooper, A. J.; Dryfe, R. a. W.; Kinloch, I. A. (2015-04-09). "How to get between the sheets: a review of recent works on the electrochemical exfoliation of graphene materials from bulk graphite". Nanoscale. 7 (16): 6944–6956. doi:10.1039/C4NR06942K. ISSN   2040-3372.
  13. 1 2 3 4 Paton, Keith R.; Varrla, Eswaraiah; Backes, Claudia; Smith, Ronan J.; Khan, Umar; O’Neill, Arlene; Boland, Conor; Lotya, Mustafa; Istrate, Oana M.; King, Paul; Higgins, Tom; Barwich, Sebastian; May, Peter; Puczkarski, Pawel; Ahmed, Iftikhar (June 2014). "Scalable production of large quantities of defect-free few-layer graphene by shear exfoliation in liquids". Nature Materials. 13 (6): 624–630. doi:10.1038/nmat3944. hdl: 2262/73941 . ISSN   1476-4660.
  14. 1 2 Parvez, Khaled; Yang, Sheng; Feng, Xinliang; Müllen, Klaus (2015-12-01). "Exfoliation of graphene via wet chemical routes". Synthetic Metals. Reviews of Current Advances in Graphene Science and Technology. 210: 123–132. doi:10.1016/j.synthmet.2015.07.014. hdl: 11858/00-001M-0000-0029-74E6-0 . ISSN   0379-6779.
  15. 1 2 3 4 5 Tao, Hengcong; Zhang, Yuqin; Gao, Yunnan; Sun, Zhenyu; Yan, Chao; Texter, John (2017-01-04). "Scalable exfoliation and dispersion of two-dimensional materials – an update". Physical Chemistry Chemical Physics. 19 (2): 921–960. doi:10.1039/C6CP06813H. ISSN   1463-9084.
  16. 1 2 Hernandez, Yenny; Nicolosi, Valeria; Lotya, Mustafa; Blighe, Fiona M.; Sun, Zhenyu; De, Sukanta; McGovern, I. T.; Holland, Brendan; Byrne, Michele; Gun'Ko, Yurii K.; Boland, John J.; Niraj, Peter; Duesberg, Georg; Krishnamurthy, Satheesh; Goodhue, Robbie (September 2008). "High-yield production of graphene by liquid-phase exfoliation of graphite". Nature Nanotechnology. 3 (9): 563–568. arXiv: 0805.2850 . doi:10.1038/nnano.2008.215. ISSN   1748-3395.
  17. 1 2 3 4 5 Zheng, Weiran; Lee, Lawrence Yoon Suk (2022-01-11). "Beyond sonication: Advanced exfoliation methods for scalable production of 2D materials". Matter. 5 (2): 515–545. doi:10.1016/j.matt.2021.12.010. ISSN   2590-2393.
  18. Xu, Yanyan; Cao, Huizhe; Xue, Yanqin; Li, Biao; Cai, Weihua (November 2018). "Liquid-Phase Exfoliation of Graphene: An Overview on Exfoliation Media, Techniques, and Challenges". Nanomaterials. 8 (11): 942. doi: 10.3390/nano8110942 . ISSN   2079-4991. PMC   6265730 . PMID   30445778.
  19. Varrla, Eswaraiah; Paton, Keith R.; Backes, Claudia; Harvey, Andrew; Smith, Ronan J.; McCauley, Joe; Coleman, Jonathan N. (2014-09-26). "Turbulence-assisted shear exfoliation of graphene using household detergent and a kitchen blender". Nanoscale. 6 (20): 11810–11819. doi:10.1039/C4NR03560G. hdl: 2262/73782 . ISSN   2040-3372.
  20. "Make graphene in your kitchen with soap and a blender". New Scientist. Retrieved 2023-11-09.
  21. 1 2 Zhang, Yi; Zhang, Luyao; Zhou, Chongwu (2013-10-15). "Review of Chemical Vapor Deposition of Graphene and Related Applications". Accounts of Chemical Research. 46 (10): 2329–2339. doi:10.1021/ar300203n. ISSN   0001-4842.
  22. 1 2 Reina, Alfonso; Jia, Xiaoting; Ho, John; Nezich, Daniel; Son, Hyungbin; Bulovic, Vladimir; Dresselhaus, Mildred S.; Kong, Jing (2009-01-14). "Large Area, Few-Layer Graphene Films on Arbitrary Substrates by Chemical Vapor Deposition". Nano Letters. 9 (1): 30–35. doi:10.1021/nl801827v. ISSN   1530-6984.
  23. 1 2 Acik, Muge; Chabal, Yves (2012-12-06). "A Review on Reducing Graphene Oxide for Band Gap Engineering". Journal of Materials Science Research. 2 (1): 101. doi: 10.5539/jmsr.v2n1p101 . ISSN   1927-0585.
  24. Su, Ching-Yuan; Lu, Ang-Yu; Xu, Yanping; Chen, Fu-Rong; Khlobystov, Andrei N.; Li, Lain-Jong (2011-03-22). "High-Quality Thin Graphene Films from Fast Electrochemical Exfoliation". ACS Nano. 5 (3): 2332–2339. doi:10.1021/nn200025p. ISSN   1936-0851.
  25. Liu, Fei; Wang, Chaojun; Sui, Xiao; Riaz, Muhammad Adil; Xu, Meiying; Wei, Li; Chen, Yuan (December 2019). "Synthesis of graphene materials by electrochemical exfoliation: Recent progress and future potential". Carbon Energy. 1 (2): 173–199. doi: 10.1002/cey2.14 . ISSN   2637-9368.
  26. 1 2 Zhang, Chen; Lv, Wei; Xie, Xiaoying; Tang, Daiming; Liu, Chang; Yang, Quan-Hong (2013-10-01). "Towards low temperature thermal exfoliation of graphite oxide for graphene production". Carbon. 62: 11–24. doi:10.1016/j.carbon.2013.05.033. ISSN   0008-6223.
  27. Choi, Wonbong; Choudhary, Nitin; Han, Gang Hee; Park, Juhong; Akinwande, Deji; Lee, Young Hee (2017-04-01). "Recent development of two-dimensional transition metal dichalcogenides and their applications". Materials Today. 20 (3): 116–130. doi: 10.1016/j.mattod.2016.10.002 . ISSN   1369-7021.
  28. Kaur, Harneet; Coleman, Jonathan N. (September 2022). "Liquid‐Phase Exfoliation of Nonlayered Non‐Van‐Der‐Waals Crystals into Nanoplatelets". Advanced Materials. 34 (35). doi:10.1002/adma.202202164. hdl: 2262/101345 . ISSN   0935-9648.