4

Nanostructured Films: Langmuir–Blodgett (LB) and Layer-by-Layer (LbL) Techniques

R.F. de Oliveira*
A. de Barros**
M. Ferreira
*    Brazilian Center for Research in Energy and Materials, Brazilian Nanotechnology National Laboratory, Campinas, São Paulo, Brazil
**    State University of Campinas, Institute of Chemistry, Campinas, São Paulo, Brazil
    Federal University of São Carlos, Center for Sciences and Technology for Sustainability, Sorocaba, São Paulo, Brazil

Abstract

This chapter discusses the historical and technological development of two important techniques for fabricating ultrathin films: Langmuir–Blodgett (LB) and layer-by-layer (LbL) methods. For the LB method, the theoretical aspects of characterizing Langmuir monolayers by pressure and surface potential are presented, and the transfer ratio and types of LB films obtained during deposition are reviewed. For the LbL technique, the physical and chemical processes involved in the film formation, the different types of interactions between materials, and the spray- and spin-assisted LbL methodologies for film production are discussed. Applications of the LB and LbL techniques are cited to illustrate their use in various fields.

Keywords

Langmuir–Blodgett
LbL
surface pressure
surface potential
transfer ratio
nanostructured films

4.1. Introduction

Ultrathin film fabrication is an important process in nanoscience and nanotechnology, as demonstrated by the large number of publications about manufacturing new thin-film materials and their applications in a wide range of fields, such as biotechnology, electronics, medicine, and so on [13]. Of the existing film fabrication methods, Langmuir–Blodgett (LB) and layer-by-layer (LbL) techniques are notable because they allow the film thickness and architecture to be controlled at the molecular level. For example, films with a high degree of molecular organization can be obtained, and the spatial arrangement of the film material can be investigated by the LB technique. The LbL method, in turn, is advantageous because of the simplicity of the film deposition process and the large variety of materials that can be used.
In addition to the technological importance of nanostructured films, fundamental questions about their formation are of great scientific interest because a wide variety of materials and film architectures are possible. The LB technique, which was developed in the mid-1900s, and the LbL method, which was discovered just over 20 years ago, are still attracting the interest of specialized research groups worldwide.
In this chapter, the key aspects of the LB and LbL techniques in ultrathin film fabrication, including the film formation mechanisms, main characterization techniques and applications, are addressed to provide the reader with an overview of these important nanotechnological processes.

4.2. Langmuir–Blodgett Technique

4.2.1. History

The first surface chemistry experiments that stimulated further research on ultrathin film formation were performed in the 18th century. According to Behroozi et al. [4], during one of his overseas journeys, Benjamin Franklin (1706–90) noted that the waves near ships with oil leakages appeared to be dampened. To investigate this phenomenon, he performed experiments on the damping effect of oil on water, which resulted in the first publication in the field in 1774. In one of these experiments, Franklin poured a small known amount of oil on a water surface and observed a decrease in the wind turbulence within a certain distance of the oil-coated area [5].
At that time, Franklin focused only on the wave-damping phenomenon and did not realize that this effect was related to the formation of a monomolecular layer at the surface. Dividing the oil volume by the area it occupies gives the oil layer thickness, which is on the nanometer scale [6]. This simple calculation was not performed until years later, when Lord Rayleigh (1842–1919) determined the thickness of oil layers. However, he did not know that the calculated value corresponded exactly to the length of the molecule used in the experiments, indicating that the layer was one molecule thick [6].
During the same period, Agnes Pockels (1862–1935) made an important advance in the science of monolayer formation by creating a prototype of the Langmuir trough (Fig. 4.1). This prototype was a primitive device in which barriers were used both to compress oil molecules scattered on a water surface and to remove impurities from the surface. Pockles also developed a method for measuring the surface tension of water in a container [5]. In recognition of these important contributions, Rayleigh helped Pockles publish her results in the prestigious journal Nature. A few years later, Irving Langmuir (1881–1957) introduced the concept of molecular conformations.
image
Figure 4.1 Schematic representation of a Langmuir trough and its accessories.
For Langmuir, molecules were asymmetrical and would therefore have identical orientations on the water surface, depending on their hydrophilic or hydrophobic character. Langmuir also estimated the size of the molecule used in his experiments, which had a large impact on the scientific community at the time. These studies motivated Langmuir to investigate monolayer formation on water surfaces, and he was awarded the Nobel Prize in Chemistry for this work in 1932. Langmuir suggested that the monolayers formed on a water surface could be transferred to a solid surface. His assistant Katherine Blodgett (1898–1979) performed the corresponding experiments, and the initial results were published in 1934 and 1935. Blodgett was awarded her own Nobel Prize for this work. During this period, the LB monolayer deposition method was developed. Several years later, Langmuir and Vincent Schaefer (1906–93) studied protein deposition on solid substrates and discovered a new approach called the Langmuir–Schaefer (LS) method for depositing Langmuir monolayers. In the LS method, a Langmuir monolayer is deposited on a surface horizontally instead of vertically, as in LB films [5,7].
The formation of Langmuir monolayers and LB and LS films was not actively studied for a long time because they had no practical application. Around 1980, research interest in this area was renewed by the use of LB films in organic electronics. Since then, physicists, chemists, biologists, and engineers have actively collaborated to develop this technique using various types of materials for a wide range of applications.

4.2.2. Description of the Technique

As illustrated in Fig. 4.1, the Langmuir trough is composed of an inert material, typically Teflon. The trough accessories include mobile barriers that compress the molecules spread at the air–water interface, a device (dipper) that immerses the substrate for the monolayer deposition, a Wilhelmy sensor that measures the surface tension of the subphase, and a capacitor that measures the surface potential.
Langmuir monolayers and LB films are fabricated by pouring a known volume of a given solution at the air–water interface. Ideally, the material in the solution is soluble in an organic solvent or solvent mixture. During film fabrication, the solvent is evaporated after a given amount of time, and the molecules are then compressed until maximum order is achieved.
Initially, the materials employed in the Langmuir technique were necessarily amphiphilic compounds, that is, molecules with a polar head group (hydrophilic) and a nonpolar tail group (hydrophobic) essentially insoluble in polar solvents. These compounds self-organize with their polar part in the water and nonpolar part in the air. The hydrophobic portion of these molecules generally consists of aliphatic chains, which reduce its solubility in the aqueous subphase. The hydrophilic part is responsible for spreading the molecules because it interacts strongly with the water. The molecular orientation of these molecules at the air–water interface minimizes their free energy [5,79].
A typical isotherm for amphiphilic molecules is illustrated in Fig. 4.2. Before compression by the barriers, the molecules are initially in the gas phase (stage A); they are fully dispersed and do not interact. As the molecules are compressed, their proximity leads to the formation of an expanded liquid phase (stage B). When the molecule surface density increases, the regular arrays in the film form, resulting in a compact structure called the condensed-liquid phase (stage C). However, further compression of the monolayer causes molecular disorder and a phenomenon commonly called collapse (stage D) [5,79].
image
Figure 4.2 Typical surface pressure isotherm for a stearic acid monolayer.
The use of LB films composed of common amphiphilic molecules is limited by their properties. To search for new technological applications of LB films, researchers have investigated the use of various materials, including polymers [1013], phospholipids [1416], enzymes [1720], and peptides [21,22]. However, the formation of monolayers of other types of molecules is not trivial because they might not be completely soluble in volatile organic solvents or might be unstable, making them difficult to spread at the surface and consequently deposit homogeneously on a substrate [8].
To minimize these difficulties, several strategies have been employed depending on the type of film material. For example, miscible solvents can be used, or the material can be spread in the subphase [10,23]. The organization, homogeneity, deposition quality, and formation of Langmuir monolayers and LB films can be evaluated experimentally by characterization methods such as Brewster angle microscopy [24,25], polarization modulation infrared reflection-absorption spectroscopy [24,26,27], and atomic force microscopy (AFM) [28,29].

4.2.3. Pressure and Surface Potential

The pressure and surface potential of LB monolayers are measured as a function of the average molecular area occupied at the air–water interface to assess molecular packing and order [5,7,9]. These characterizations are fundamental and necessary for evaluating the quality of the Langmuir film formed at the interface. To determine the surface pressure of a Langmuir monolayer, the material is spread on an aqueous subphase, and the surface pressure is measured by detecting variations in the liquid surface tension, as shown by the following equation:

π=γ0γA

image(4.1)
where γ0 and γA are the surface tensions of pure water and due to the monolayer, respectively [5,7].
The surface pressure isotherms are obtained using a Wilhelmy sensor, which is suspended by a wire from an electrobalance and partially immersed in the water. The electrobalance measures the force exerted to keep the sensor stationary as the surface tension varies. The vertical force exerted by the surface tension is detected by the balance and converted into a voltage [5,7]. A typical isotherm of an amphiphilic material (stearic acid) is illustrated in Fig. 4.2.
One disadvantage of this technique is related to the contact angle (θ) formed between the liquid subphase and the balance sensor, as shown in Fig. 4.3 [5,7]. Another disadvantage is that the position of the Wilhelmy sensor relative to the barrier positions can affect the pressure isotherm during monolayer formation. These problems can be minimized by placing the sensor in the center of the trough [8].
image
Figure 4.3 Schematic representation of the contact angle between the Wilhelmy sensor and the subphase.
The surface pressure isotherms might exhibit hysteresis. This phenomenon arises due to differences in the molecular behavior during compression and decompression. Specifically, as the barriers compress the molecules, the area that they occupy decreases, resulting in an increase in their interactions and consequently an increase in the surface pressure. When the monolayer is decompressed, the surface pressure gradually decreases, but not in the same way that it increases during the initial compression. The hysteresis in the area occupied by the molecules indicates that some of the material is lost to the aqueous subphase or to aggregation. A large hysteresis implies that the stability of the monolayer is low. Another method for characterizing monolayer stability is based on compressing the monolayer to the maximum pressure at which molecule order is observed for long periods of time. If large variations in the area are observed during this period, the monolayers are unstable and therefore unsuitable for film deposition [79].
Another widely used technique for characterizing Langmuir films is surface potential (∆V) measurements. Surface potential is defined as the difference in the potentials of the aqueous subphase covered with the monolayer and the pure aqueous subphase. Surface potential can be measured using a Kelvin probe or a vibrating plate capacitor, which is more common [5,79,30]. In the vibrating capacitor method, surface potential is measured by a plate positioned above the water that detects the vibrations of the spread molecules. The potential of the pure water is used as a reference and is measured by a metal plate immersed in the subphase. The ∆V value is the change in the permanent electric dipoles at the air–water interface due the presence of a partially or fully ionized film. To gain insight into the effect, theoretical models were developed to relate the measured potential to the dipole moments of the film material. The most commonly used model was developed by Demchak and Fort (DF) [30]. In this model, the Langmuir monolayer is considered to be a three-layer capacitor in which each layer has a different relative permittivity, as illustrated in Fig. 4.4 [8,9,30].
image
Figure 4.4 Capacitor model of Demchak and Fort for a Langmuir monolayer. Adapted from P. Dynarowicz-Latka, A. Dhanabalan, O.N. Oliveira Jr., Modern physicochemical research on Langmuir monolayers, Adv. Colloid Interface Sci. 91 (2) (2001) 221–293 [8].
According to the DF model, the surface potential is given by:

V=1Aɛ0μ1ɛ1+μ2ɛ2+μ3ɛ3+Ψ0

image(4.2)
where A is the average area per molecule; ɛ0 is the vacuum permittivity; μ1, μ2, and μ3 are the dipole moments attributed to the polarization and reorientation of the molecules; and ɛ1, ɛ2, and ɛ3 are the permittivities due to changes in the dipole moments of the hydrophilic and hydrophobic groups. More specifically, the component μ1/ɛ1 corresponds to the reorientation of water molecules that is induced by the presence of the monolayer, μ2/ɛ2 is the contribution of the hydrophilic portion of the molecule, μ3/ɛ3 is the contribution of the hydrophobic portion, and Ψ0 is attributed to the electrical double layer that forms when the film is partially or fully ionized [5,79]. In polymer films, for example, it is impossible to estimate values for μ1/ɛ1 + μ2/ɛ2 + μ3/ɛ3 and Ψ0, mainly because the polymer is often processed under different conditions, which induce structural changes in the polymer chain. These changes make it difficult to explain the experimental results using a theoretical surface potential model.
Molecular behavior during the surface potential measurement can be described as follows. Initially, the molecules spread on the subphase cover an area so large that their interactions are too weak to induce a detectable change in the surface potential of the aqueous subphase. During compression, the area covered by the molecules reaches a critical value at which the potential is no longer zero, and it increases sharply as the area per molecule is further decreased. Due to this critical area, surface potential measurements are more sensitive to the film organization than surface pressure measurements [5,79]. Fig. 4.5 shows a typical surface potential curve for stearic acid based on the DF model, which considers the changes in the dielectric constants of the three layers as functions of the area per molecule.
image
Figure 4.5 Typical surface potential isotherm for a stearic acid monolayer.

4.2.4. Langmuir Monolayer Deposition: Transfer Ratio

Langmuir monolayers can be deposited by two methods, the vertical (LB) and horizontal (LS) methods, which are as illustrated in Fig. 4.6. The vertical method is more commonly used because the amount of material deposited on the substrate cannot be effectively controlled during horizontal deposition, that is, it is unknown if the entire surface is actually covered by the monolayer [5,31]. The effectiveness of the monolayer deposition process is described by the transfer ratio (TR). A TR of close to 1.0 implies that the material exhibits good adhesion to the substrate when it is immersed in and then removed from the subphase. However, a low TR value indicates that the material does not adhere to the substrate well, and consequently, the second monolayer deposited on the film might be easily removed, resulting in a low-quality film [7]. The TR value is calculated using the following equation:
image
Figure 4.6 Schematic representation of the different Langmuir monolayer deposition methods.
(A) Vertical deposition (LB) and (B) horizontal deposition (LS).

τ=ALAS

image(4.3)
where AL is the decrease in the area occupied by the molecules at the air/water interface (at constant pressure) and AS is the area of the substrate covered with the monolayer [7].
Different LB film architectures can be obtained by varying the deposition parameters, substrate characteristics, and film material. The possible architectures are called X-, Y-, and Z-type films and are illustrated in Fig. 4.7. Y-type LB films are generally obtained when substrates with hydrophilic surfaces are used. For the X- and Z-type films, deposition occurs preferentially during substrate immersion and removal from the subphase, which favors the formation of films in which the polar and nonpolar groups in adjacent monolayers interact. It is important to note that Z-type films form on hydrophilic substrates, whereas X-type films form onto hydrophobic substrates [5,7,32].
image
Figure 4.7 Different LB film types: X (A), Y (B), and Z (C).

4.2.5. Applications

Numerous applications of LB and LS films, such as nonlinear optical and piezoelectric devices [33,34], chemical sensors and biosensors [10,12,13,1720], cell membrane applications [3537], and so on [1,3840] have been reported in the literature.
LB and LS films are widely used in sensors, biosensors, and cell membrane model studies because their molecular organization can be highly controlled, which results in satisfactory experimental results. In one study, enzymes were immobilized on lipid monolayers of LS films, and the catalytic activity of the biomolecule-containing films in sucrose hydrolysis was evaluated. The experimental results indicated that 78% of the invertase enzymatic activity was maintained in this system [17]. In another study, the interactions between mucin and chitosan were studied using dimyristoyl phosphatidic acid (DMPA) Langmuir monolayers as cell membrane models. The experimental results for the DMPA–chitosan–mucin Langmuir and LB films indicated that a mucin–chitosan complex forms via electrostatic interactions that are crucial for the mucoadhesion mechanism [35].
De Barros et al. developed electrochemical sensors composed of polyaniline and montmorillonite clay using LB technique for the detection of copper, lead, and cadmium metal ions. The Langmuir isotherms and ultraviolet-visible (UV-vis) absorption spectroscopy, Fourier transform infrared spectroscopy and Raman spectroscopy results indicated the existence of synergistic molecular-level interactions between the film materials, which were favorable for the sensor performance, that is, the sensors had high sensitivity and a low detection limit on the order of μg L−1 [10]. Thin films of polythiophene derivatives were fabricated using LS and spin-coating techniques to compare the effect of the deposition method on the electrical properties of the films. The experimental results revealed an electrochromic effect during analysis, and the electrical and electrochemical measurements revealed the effects of the film organization on its electrical properties [11].
For further reading on the theoretical concepts and applicability of this technique, the reader is referred to the work of M. Petty titled “Langmuir–Blodgett Films: An Introduction” and other texts in this field.

4.3. Layer-by-Layer Technique

4.3.1. History

The LbL method became widely spread during the 1990s due to several publications by G. Decher [4146], including a paper in the prestigious journal Science [46] in 1997. This technique, however, was first described in 1966 by R.K. Iler as a study titled “Multilayers of Colloidal Particles” [47]. This study showed that multilayer films could be obtained by alternately immobilizing oppositely charged colloidal particles. Iler observed that the thicknesses of alternating cationic (alumina) and anionic (silica) layers on glass could be controlled and observed by the change in color of light reflected on the substrate surface [47].
The LbL technique was further developed in several subsequent studies that described, for example, the successive deposition of inorganic ionic compounds [48] and polyanion adsorption on cellulosic fibers containing preadsorbed polycations [49], among other systems [50]. Decher contributed not only to the rediscovery of this technique but also to its use as an alternative to the LB method for fabricating nanostructured films. He also demonstrated that it could be used to deposit films of various types of materials, including bipolar amphiphilic molecules [42], polyelectrolytes [42,44], and proteins and DNA [43,44]. According to Decher and Schlenoff [50], the field of nanostructure fabrication via the LbL method is still popular and growing more than 20 years after the technique was reintroduced.

4.3.2. Description of the Technique

As the terminology itself suggests, the LbL technique is a method for obtaining nanostructured films by successively depositing layers of different materials, such as polymers, nanoparticles, enzymes, cells, and so on, with highly controlled thicknesses at the molecular level. The ultrathin films fabricated using LbL method have nanostructures that can be used in optical devices, electronics, sensors, and biotechnological applications [2,3,51,52].
Unlike the sophisticated LB method, the main advantages of the LbL technique are its simplicity and low cost. Fig. 4.8 illustrates the LbL deposition process and film formation.
image
Figure 4.8 Schematic representation of the traditional LbL deposition process.
Initially, a substrate is immersed in a solution containing charged species, for example, polyanions (step A). During this step, the anionic polyelectrolyte adsorbs on the substrate surface via simple electrostatic interactions, resulting in the formation of a negative charge network on the surface. Then, the substrate is immersed in a rinsing solution (step B) to remove weakly adsorbed material to prevent contamination of the next solution. The substrate is subsequently immersed in a polycation solution to generate a new network of positive charges on the surface (step C). Finally, the substrate is again immersed in a rinsing solution (step D). Thus, a bilayer of the ionic materials, in this case polyelectrolytes, is obtained. This procedure can be repeated as many times as desired to obtain multilayer films with controlled structures and thicknesses.

4.3.3. Mechanisms of LbL Film Formation

The multilayer formation process can be divided into two stages: (1) the initial adsorption of the material on the surface (fast process) and (2) the relaxation of the adsorbed layer (slower process) [53]. In the first stage, the polyelectrolyte segments diffuse through the solution to adsorb on the surface until a repulsive electrical potential is generated, preventing the adsorption of more chains [54]. At this point, the polyelectrolyte layer has reached saturation, which is only possible when the adsorption rate is not limited, that is, when the species concentration in solution is much larger than the saturation concentration of the adsorbed species [46,53,54]. In this stage, the polyelectrolyte segments adsorb on the surface because of favorable electrostatic interactions, and the surface becomes oppositely charged, which enables the adsorption of new polyions with opposite charges to the surface [46].
When the initial substrate charge density is small, multiplication of surface functionality occurs. In this process, the charge density of the first adsorbed layer is larger than that of the substrate, which favors the subsequent adsorption of oppositely charged species [46]. It should be noted that the surface charge and roughness of the substrate (glass, quartz, silicon, paper, cloth, etc.) might affect the deposition of the first few layers [53]. The influence of the substrate on the deposited layers becomes negligible after a few deposition cycles. Then, the film growth and structure are fundamentally governed by the film materials. This important feature allows a film to be fabricated on a different surface for a given application than on the substrate that was used to study the film properties [50].
During the second stage of the layer formation process, the conformations of the adsorbed polyelectrolyte chains change as they relax [54]. The relaxation process explains the interpenetration of adjacent polyelectrolyte chains observed in experimental studies. However, it is assumed that the adsorbed segments do not diffuse along the surface during LbL film formation because this process would involve the breaking of many electrostatic bonds simultaneously, which requires more energy than that gained by thermodynamic contributions to the system and is therefore highly improbable [54].
Although electrostatic interactions are involved in the formation of polyelectrolyte films (Fig. 4.8), they are not the only interactions that can be exploited in LbL film formation. Depending on the characteristics of the selected materials, covalent bonds, van der Waals forces, and hydrogen bonds can also govern the film formation mechanism, further enhancing the versatility of the LbL technique [55].
Hydrogen bonds, for example, are sensitive to the medium and can be broken and reformed by changing the pH. This phenomenon is observed in the production of some porous films, such as polyacid acrylic (PAA)/polyvinyl pyridine (PVP) multilayers. In a basic solution, the PAA layers dissolve, allowing the PVP layers to subsequently rearrange to yield the porous structure [56]. The film characteristics, such as the pore size and pore size distribution, can also be tuned by varying the pH and temperature of the solution and the immersion time and nature of the substrate during treatment in a basic solution [55,57]. Hydrogen bonds also enable the production of multilayers of a single material; for example, films can be fabricated from a dendrimer with carboxylic acid groups that can act as both hydrogen bond donors and acceptors [56]. Other examples of different multilayer film formation mechanisms include charge transfer reactions (electron donor and acceptor materials are alternately deposited) [58] and covalent bonding [59].
In addition to these different interactions, other strategies, such as using avidin–biotin biospecificity [60], DNA hybridization [61], and other interactions [2,55,62,63], have been employed to obtain LbL films. In LbL films, different interactions not only drive the film formation mechanism but also participate to a minor degree in films formed via electrostatic interactions, influencing their properties, such as their stability, morphology, and thickness [53].
Of all the characteristics of a film, its thickness might be one of the most important [50]. The thickness depends on the deposition conditions of each layer, which in turn influence the subsequently deposited layer and, consequently, the final film properties, including the roughness, uniformity, chemical resistance, and so on. The intrinsic film material properties, such as the nature and density of the charged groups; the solution characteristics, such as the concentration, pH, and ionic strength; the fabrication operating parameters, such as the deposition and rinse times; and many other factors affect the film formation. For example, studies indicate that in general, increasing the ionic strength of a polyelectrolyte solution causes the polymer chains to contract, which leads to an increase in the surface density of the adsorbed segments and thus to an increase in film thickness [54]. However, exposing a polyelectrolyte film to a concentrated salt solution after fabrication leads to a decrease in film roughness due to increased chain mobility [50]. Moreover, changing the solution pH can alter the degree of dissociation and the conformation of some polyelectrolytes and also the enzymatic activity of biological films. Although many factors in the deposition process affect the film properties, understanding their effects are crucial to optimizing the film deposition process.

4.3.4. Spray- and Spin-LbL Methods

In addition to the conventional immersion method, other LbL film fabrication procedures have been reported in the literature. In the spin-assisted LbL method, the film materials are alternately added to a substrate rotating at a given speed, whereas in the spray-LbL method, the materials are sprayed in small liquid particles on the substrate surface. These methods are depicted in Fig. 4.9.
image
Figure 4.9 Novel LbL deposition strategies.
(A) Spin-assisted LbL and (B) spray-LbL.
These techniques can be advantageous over the conventional method and therefore more appropriate, depending on the desired film characteristics and application. The spin-assisted LbL and spray-LbL methods are faster than the conventional immersion method, particularly in the production of thick (micrometer) films because they do not depend on the diffusion kinetics of the species in the liquid medium. They also consume smaller amounts of material. However, the spin-assisted LbL method can only be used with flat substrates and produces nonuniform films on large-area substrates. The spray technique is more appropriate for these types of substrates [64]. The characteristics of films produced via the spray and conventional methods are very similar, whereas films produced by the spin-assisted LbL method are generally smoother due to decreased chain interpenetration, for example, in the case of polyelectrolytes [64]. It should be noted that all three deposition methods can be automated, which increases the reproducibility of the fabricated films [65].

4.3.5. Characterization Methods

Regardless of the deposition procedure, materials, and desired application, the characterization of LbL films is frequently necessary. In particular, UV-vis spectroscopy, zeta potential determination, and quartz crystal microbalances are commonly used to monitor the film during the deposition steps. For example, the amount of material adsorbed in each step can be determined by UV-vis monitoring in which the absorption band intensities of compounds with chromophore groups are measured as a function of the number of deposited layers. However, this and some other techniques require the film to be dried after rinsing, which might not be desirable depending on the application. Alternatively, film growth can be monitored in situ by a quartz crystal microbalance. In this method, the frequency of a resonant quartz crystal changes in response to changes in the material mass, enabling the amount of material adsorbed during each step to be measured. This technique allows film growth to be monitored when UV-vis spectroscopy is not possible, for example, when the material of interest absorbs only at frequencies outside the ultraviolet and visible regions [50].
Several other techniques, such as infrared spectroscopy, Raman spectroscopy, AFM, scanning electron microscopy (SEM), and transmission electron microscopy (TEM), are often used to characterize LbL films. Raman and infrared spectroscopies are used to determine the functional groups present in the film materials by recording their vibrational modes. Techniques such as AFM, SEM, and TEM are used to analyze film morphology, including its surface roughness, uniformity, and so on. The reader is referred to other chapters in this book for more information on various nanostructure characterization techniques.

4.3.6. Applications

Although the LbL technique has been widely used to fabricate polyelectrolyte films, it is currently employed for many different materials and applications. A few of the commonly used synthetic and natural materials include polypeptides, proteins, enzymes, polysaccharides, metallic nanoparticles, oxides and clays, DNA, biopolymers, dendrimers, and carbon nanotubes [2,50,6671]. LbL films are employed in biotechnological applications, such as antibacterial and antiadhesive biopolymer surface coatings [72], sensors [66,71] and electrochemical biosensors [7375], and also in biomedical applications, such as self-supporting polymeric shells for controlled drug delivery [53]. In the latter application, polystyrene, polylactic acid and silica particles are coated with polyelectrolyte or biopolymer multilayers (steps I and II, Fig. 4.10) and then decomposed in an appropriate solvent to obtain spherical shells (step III, Fig. 4.10). The drug of interest is introduced into the shells via a stimulus, such as a change in the temperature, magnetic field, or pH, and then released at the target using the same stimulus [53]. Fig. 4.10 illustrates the formation process of these spherical shells.
image
Figure 4.10 Schematic illustration of the fabrication of spherical shells by the LbL method.
LbL films are also used in electronic devices; they serve as charge injection layers in light-emitting diodes [76], are employed in solar cells [77], constitute the active layer in field-effect transistors [78], and are used in memory applications [79]. In addition, these films are used as ion-exchange membranes in fuel cells [80] and as antireflective surface coatings in optical applications [81], among other applications [2,3,9].
For a more comprehensive study of the concepts and applicability of the LbL technique, the reader is referred to the publication of G. Decher titled “Multilayer Thin Films: Sequential Assembly of Nanocomposite Materials, 2nd Edition” and other texts in this field.

4.4. Final Considerations

The unique properties of nanostructured films constructed from various materials and the versatility of the LB and LbL methods have contributed considerably to the development of nanotechnology. LB and LbL methods have enabled the use of various materials with diverse architectures in many technological applications, from electronics to medicine. It should be noted that these methods have limitations and are thus complementary. Although ultrathin film fabrication has been extensively studied, it is a growing and popular field with many possibilities yet to be explored.

Acknowledgment

The authors thank Bianca Martins Estevão for producing some of the figures.

References

[1] Ariga K, et al. 25th anniversary article: what can be done with the Langmuir–Blodgett method? Recent developments and its critical role in materials science. Adv. Mater. 2013;25(45):64776512.

[2] Ariga K, et al. Layer-by-layer nanoarchitectonics: invention, innovation and evolution. Chem. Lett. 2014;43(1):3668.

[3] Ariga K, et al. Thin film nanoarchitectonics. J. Inorg. Organomet. Polym. Mater. 2015;25(3):466479.

[4] Behroozi P, et al. The calming effect of oil on water. Am. J. Phys. 2007;75:407414.

[5] A. Ulman, Langmuir–Blodgett films, in: An Introduction to Ultrathin Organic Films from Langmuir–Blodgett to Self-Assembly, [s.I.] Academic Press, USA, 1991, pp. 101–219.

[6] J.N. Israelachvili, Historical perspective, in: Intermolecular and Surface Forces, third ed., [s.I] Academic Press, USA, 2011, pp. 15–16.

[7] Petty MC. Langmuir–Blodgett films: an introduction [s.I.]. Cambridge, England: Cambridge University Press; 1996.

[8] Dynarowicz-Latka P, Dhanabalan A, Oliveira Jr ON. Modern physicochemical research on Langmuir monolayers. Adv. Colloid. Interface Sci. 2001;91(2):221293.

[9] O.N. Oliveira Jr., F.J. Pavinatto, D.T. Balogh, Fundamentals and applications of organised molecular films, in: Nanomaterials and Nanoarchitectures. first ed., [s.I] Springer, 2015, pp. 301–343.

[10] De Barros A, et al. Synergy between polyaniline and OMt clay mineral in Langmuir–Blodgett films for the simultaneous detection of traces of metal ions. ACS Appl. Mater. Interfaces. 2015;7(12):68286834.

[11] Braunger ML, et al. Electrical and electrochemical measurements in nanostructures films of polythiophenes derivates. Electrochim. Acta. 2015;164(1–6).

[12] Chen X, et al. Enhanced of polymer solar cells with a monolayer of assembled gold nanoparticle films fabricated Langmuir–Blodgett technique. Mater. Sci. Eng. B. 2013;178(1):5359.

[13] Jayaraman S, Yu LT, Srinivasan MP. Polythiphene-gold nanoparticle hybrid systems: Langmuir–Blodgett assembly of nanostructure films. Nanoscale. 2013;5:29742982.

[14] Guzmán E, et al. DPPC-DOPC Langmuir monolayers modified by hydrophilic silica nanoparticles: phase behaviour, structure and rheology. Colloids Surf. A Physicochem. Eng. Asp. 2012;413:174183.

[15] Nakahara H, et al. Interfacial properties in Langmuir monolayers and LB films of DPPC with partially fluorinated alcohol (F8H7OH). J. Oleo Sci. 2013;62(12):10171027.

[16] Saha M, Hussain SA, Bhattacharjee D. Interaction of nano-clay platelets with a phospholipid in presence of a fluorescence probe. Mol. Cryst. Liq. Cryst. 2015;608(1):198210.

[17] Cabaj J, et al. Biosensing invertase-based Langmuir–Schaefer films: preparation and characteristic. Sens. Actuators. B. 2012;166/167:7582.

[18] Caseli L, Siqueira JR. High enzymatic activity preservation with carbon nanotubes incorporated in urease-lipid hybrid Langmuir–Blodgett films. Langmuir. 2012;28(12):53985403.

[19] Rocha JM, Caseli L. Adsorption and enzyme activity of sucrose phosphorylase on lipid Langmuir and Langmuir–Blodgett films. Colloids Surf. B. 2014;116:497501.

[20] Scholl FA, Caseli L. Langmuir and Langmuir–Blodgett films of lipids and penicillinase: studies on adsorption and enzymatic activity. Colloids Surf. B. 2015;126:232236.

[21] Togashi K, et al. Fabrication of Langmuir–Blodgett films using amphiphilic peptides. J Nanosci. Nanotechnol. 2012;12(1):568572.

[22] Kato N, Sasaki T, Mukai Y. Partially induced transition from horizontal to vertical orientation of helical peptides at the air-water interface and the structure of their monolayers transferred on the solid substrates. Biochim. Byophys. Acta. 2015;1848(4):967975.

[23] Hussain SA, et al. Incorporation of nano-clay saponite layers in the organo-clay hybrid films using anionic amphiphile stearic acid by Langmuir–Blodgett technique. Thin Solid Films. 2013;536:261268.

[24] Ballesteros L, et al. Directionally oriented LB films of and OPE derivate: assembly, characterization and electrical properties. Langmuir. 2011;27(7):36003610.

[25] Kucuk AC, Matsui J, Miyashita T. Langmuir–Blodgett films composed of amphiphilic double-decker shaped polyhedral oligomeric silsesquioxanes. J. Colloid Interface Sci. mar 2011;355(1):106114.

[26] Shultz MJ, et al. Aqueous solution/air interfaces probed with sum frequency generation spectroscopy. J. Phys. Chem. B. 2002;106:53135324.

[27] Blaudez D, et al. Polarization-modulated FT-IR spectroscopy of a spread monolayer at the air/water interface. Appl. Spectrosc. 1993;47:869874.

[28] Talu S, Patra N, Salerno N. Micromorphological characterization of polymer-oxide nanocomposite thin films by atomic force microscopy and fractal geometry analysis. Prog. Org. Coat. 2015;89:5056.

[29] Makowiecki J, et al. Molecular organization of perylene derivatives in Langmuir–Blodgett multilayers. Opt. Mater. 2015;46:555560.

[30] Demchak RJ, Fort Jr T. Surface dipole moments of close-packed un-ionized monolayers at the air-water interface. J. Colloid Interface Sci. 1974;46(2):191202.

[31] Y.S. Lee, Nanostructures films, in: Self Assembly and Nanotechnology, [s.I.] John Wiley & Sons, Inc., Hoboken, New Jersey, 2008, pp. 259–262.

[32] G. Cao, Two-dimensional nanostructures: thin films, in: Nanostructures & Nanomaterials: Synthesis, Properties & Applications, [s.I.] Imperial College Press, United States, 2004, pp. 173–224.

[33] Giancane G, et al. Investigations and application in piezoelectric phenol sensor of Langmuir–Schaefer films of a copper phthalocyanine derivate functionalized with bulky substituents. J. Colloid. Interface Sci. 2012;377:176183.

[34] Fernández R, et al. Optical storage in azobenzene –containing epoxy polymers process as Langmuir–Blodgett films. Mater. Sci. Eng. C. 2013;33(3):14031408.

[35] Silva CA, et al. Interaction of chitosan and mucin in a biomembrane model environment. J. Colloid Interface Sci. 2012;376(1):289295.

[36] Nobre MT, et al. Interaction of bioactive molecules and nanomaterials with Langmuir monolayers as cell membrane models. Thin Solid Films. 2015;593:158178.

[37] Mangiarotti A, Caruso B, Wilke N. Phase coexistence in films composed of DLPC and DPPC: a comparison between different model membrane system. Biochim. Biophys. Acta. 2014;1838(7):18231831.

[38] Arslanov V, et al. Designed and evaluation of sensory system based on amphiphilic anthraquinones molecular receptors. Colloids Surf. A. 2015;483:193203.

[39] Yin Z, et al. Real-time DNA detection using Pt nanoparticle-decorated reduced graphene oxide field-effect transistors. Nanoscale. 2012;4:293297.

[40] Wang L, et al. A new strategy for enhancing electrochemical sensing from MWCNTs modified electrode with Langmuir–Blodgett film and used in determination of methylparaben. Sens. and Actuators B. 2015;211:332338.

[41] Decher G, Hong J. Buildup of ultrathin multilayer films by a self-assembly process: II. Consecutive adsorption of anionic and cationic bipolar amphiphiles and polyelectrolytes on charged surfaces. Phys. Chem. Chem. Phys. 1991;95(11):14301434.

[42] Decher G, Hong JD, Schimitt J. Buildup of ultrathin multilayer films by self-assembly process: III. Consecutively alternating adsorption of anionic and cationic polyelectrolytes on charged surfaces. Thin Solid Films. 1992;210:831835.

[43] Lvov Y, Decher G, Sukhorukov G. Assembly of thin films by means of successive deposition of alternate layers of DNA and poly(allylamine). Macromolecules. 1993;26(20):53965399.

[44] Decher G, Lvov Y, Shimitt J. Proof of multilayer structural organization in self-assembled polycation-polyanion molecular films. Thin Solid Films. 1994;244(1):772777.

[45] Decher G, et al. New nanocomposite films for biosensors: layer-by-layer adsorbed films of polyelectrolytes, proteins or DNA. Biosens. Bioelectron. 1994;9(9):677684.

[46] Decher G. Fuzzy nanoassemblies: toward layered polymeric multicomposites. Science. 1997;277(5330):12321237.

[47] Iler RK. Multilayers of colloidal particles. J. Colloid Interface Sci. 1996;21(6):569594.

[48] Nicolau YF. Solution deposition of thin solid compound films by a successive ionic-layer adsorption and reaction process. Appl. Surf. Sci. 1985;22:10611074.

[49] Aksberg R, Ödberg L. Adsorption of anionic polyacrylamide on cellulosic fibers with pre-adsorbed polyelectrolytes. Nord. Pulp. Pap. Res. J. 1990;5:168171.

[50] G. Decher, F.F. Schlenoff, Multilayer Thin Films: Sequential Assembly of Nanocomposite Materials, second ed., John Wiley & Sons, Germany, 2012.

[51] Decher G, Layered nanoarchitectures via directed assembly of anionic and cationic molecules. Sauvage JP, Hosseini MW, eds. Comprehensive supramolecular chemistry [s.I.], 9. Oxford: Pergamon Press; 1996:507528.

[52] El-Khouri RJ, et al. Multifunctional layer-by-layer architectures for biological applications. In: Knoll W, Advincula RC, eds. Functional polymer films [s.I.]. Germany: Wiley-VCH; 2011:1172.

[53] De Villiers MM, et al. Introduction to nanocoatings produced by layer-by-layer (LbL) self-assembly. Adv. Drug Deliv. Rev. 2011;63(9):701715.

[54] Schönhoff M. Layered polyelectrolyte complexes: physics of formation and molecular properties. J. Phys. Condens. Matter. 2003;17811808.

[55] Borges J, Mano jF. Molecular interactions driving the layer-by-layer assembly of multilayers. Chem. Rev. 2014;114(18):88838942.

[56] Bai S, et al. Hydrogen-bonding-directed layer-by-layer polymer films: substrate effect on the microporous morphology variation. Eur. Polym. J. 2006;42(4):900907.

[57] Zhang X, Chen H, Zhang H. Layer-by-layer assembly: from conventional to unconventional methods. Chem. Commun. 2007;14:1395.

[58] Zhang J, et al. Layer-by-layer assembly of azulene-based supra-amphiphiles: reversible encapsulation of organic molecules in water by charge-transfer interaction. Langmuir. 2013;29:63486353.

[59] Broderick AH, Manna U, Lynn DM. Covalent layer-by-layer assembly of water-permeable and water-impermeable polymer multilayers on highly water-soluble and water-sensitive substrates. Chem. Mater. 2012;24:17861795.

[60] Takahashi S, Sato K, Anzai J-I. Layer-by-layer construction of protein architectures through avidin–biotin and lectin–sugar interactions for biosensor applications. Anal. Bioanal. Chem. 2012;402:17491758.

[61] Lee L, et al. Influence of salt concentration on the assembly of DNA multilayer films. Langmuir. 2010;26(5):34153422.

[62] Matsusaki M, et al. Layer-by-layer assembly through weak interactions and their biomedical applications. Adv. Mater. 2012;24(4):454474.

[63] Xu H, Schönhoff M, Zhang X. Unconventional layer-by-layer assembly: surface molecular imprinting and its applications. Small. 2012;8(4):517523.

[64] Li Y, Wang X, Sun J. Layer-by-layer assembly for rapid fabrication of thick polymeric films. Chem. Soc. Rev. 2012;41(18):5998.

[65] Richardson JJ, Bjornmalm M, Caruso F. Technology-driven layer-by-layer assembly of nanofilms. Science. 2015;348:411423.

[66] Kim Y, et al. Stretchable nanoparticle conductors with self-organized conductive pathways. Nat. Lett. 2013;500:5964.

[67] Xi Q, et al. Gold nanoparticle-embedded porous graphene thin film fabricated via layer-by-layer self-assembly and subsequent thermal annealing for electrochemical sensing. Langmuir. 2012;28:98859892.

[68] De Barros A, et al. Nanocomposites based on LbL films of polyaniline and sodium montmorillonite clay. Synth. Met. 2014;197:119125.

[69] Sham AYW, Notley SM. Graphene polyelectrolyte multilayer film formation driven by hydrogen bonding. J. Colloid Interface Sci. 2015;456:3241.

[70] Xu X, et al. Multifunctional drug carriers comprised of mesoporous silica nanoparticles and polyamidoamine dendrimers based on layer-by-layer assembly. Mater. Des. 2015;88:11271133.

[71] Silva JS, et al. Layer-by-layer films based on carbon nanotubes and polyaniline for detecting 2-chlorophenol. J. Nanosci. Nanotechnol. 2014;14(9):65866592.

[72] Follmann HDM, et al. Anti-adhesive and antibacterial multilayer films via layer-by-layer assembly of TMC/heparin complexes. Biomacromolecules. 2012;13:37113722.

[73] De Oliveira RF, et al. Exploiting cascade reactions in bienzyme layer-by-layer films. J. Phys. Chem. C. 2011;115(39):1913619140.

[74] Campos PP, et al. Amperometric detection of lactose using β-galactosidase immobilized in layer-by-layer films. ACS Appl. Mater. Interfaces. 2015;6(14):1165711664.

[75] Graça JS, et al. Amperometric glucose biosensor based on layer-by-layer films of microperoxidase-11 and liposome-encapsulated glucose oxidase. Bioelectrochemistry. 2014;96:3742.

[76] Ho PKH, et al. Molecular-scale interface engineering for polymer light-emitting diodes. Nature. mar 2000;404:481484.

[77] Saha SK, Guchhait A, Pal AJ. Organic/inorganic hybrid pn-junction between copper phthalocyanine and CdSe quantum dot layers as solar cells. J. Appl. Phys. 2012;112(4):044507.

[78] Hwang H, et al. Highly tunable charge transportation in layer-by-layer assembled graphene transistor. ACS Nano. 2012;6(3):24322440.

[79] Koo B, Baek H, Cho J. Control over memory performance of layer-by-layer assembled metal phthalocyanine multilayers via molecular-level manipulation. Chem. Mater. 2012;24(6):10911099.

[80] Zhao L, et al. Fabrication of ultrahigh hydrogen barrier polyethyleneimine/graphene oxide films by LbL assembly fine-tuned with electric field application. Compos. Part A Appl. S. 2015;78:6069.

[81] Katagiri K, et al. Anti-reflective coatings prepared via layer-by-layer assembly of mesoporous silica nanoparticles and polyelectrolytes. Polym. J. 2015;47:190194.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset