Next Article in Journal
Effect of Nozzle Diameter on Basalt Continuous Fiber Properties
Previous Article in Journal
Composite Performance Evaluation of Basalt Textile-Reinforced Geopolymer Mortar
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Probing of Nanoscale Friction and Mechanical Characteristics of Cotton Fiber’s Surface

Department of Biological and Agricultural Engineering, Texas A&M University, College Station, TX 77843, USA
*
Author to whom correspondence should be addressed.
Submission received: 7 May 2019 / Revised: 2 July 2019 / Accepted: 9 July 2019 / Published: 15 July 2019

Abstract

:
The surface topography and nanomechanical attributes of two samples of cotton fibers, namely, A and B, were characterized with various operation modes of an Atomic Force Microscope (AFM). The surface topography and friction images of the fibers were obtained in contact mode. The nanomechanical properties images—i.e., adhesion and deformation—were obtained in force tapping mode. The results indicate that the surface nanomechanical and nanoscale frictional properties of the fibers vary significantly between two samples. The plots of friction versus normal force of the fibers’ surface from both samples are fitted to the equation of single-asperity, adhesion-controlled friction. Nevertheless, within the range of the applied normal force, the friction curves of sample A surfaces show a characteristic transition phase. That is, under low normal forces, the friction curves closely conform with the Hertzian component of friction; after the transition takes place at higher normal forces, the friction curves follow Amontons’ law of friction. We demonstrated that the transition phase corresponds to a state at which the cuticle layer molecules are displaced from the fibers’ surface. The average adhesion force of the samples is consistent with the average friction signal strength collected under low normal forces.

1. Introduction

An Atomic Force Microscope (AFM) is a widely-used instrument to characterize the morphological, nanotribological, and nanomechanical properties of various surfaces [1]. Applications of AFM in the nanoscale surface characterization of plant cells are expanding [2]. Plant cells are naturally coated with an ultrathin extracellular membrane (less than hundreds of nanometers thick), collectively referred to as the cuticle layer. The main functions of the cuticle layer are to preserve the physiological integrity of the cell and strengthen its overall structural stability [3,4,5]. The cuticle layer can be regarded as a multilayer, natural coating, primarily composed of biopolyester cutin, pectins, and lipid-derived compounds.
The surface nanomechanical properties of the cuticle layer are of significant biological and technological importance. For instance, in the field of plant cell morphogenesis, it has been shown that the nanomechanical properties of the cuticle layer influence the cell’s shape changes and the growth rate [6,7,8]. In biomimicry, advances in the synthesis of artificial superhydrophobic surfaces—which mimic the hierarchical microstructure of plant leaves and petal surfaces—rely on an accurate characterization of the surface nanomechanical and nanotopographical features of plants’ cuticle layers [9]. In the industrial-scale agriculture production, the nanoscale interfacial frictional properties of natural fibers restrain their mass flow and motion throughout different fiber processing stages [10,11].
Various AFM operation modes have been developed to investigate different physical and dimensional properties of thin films at the nanoscale. In conventional contact mode, the lateral deflection of a cantilever—pertaining to the plane of a sample surface—corresponds to the nanoscale friction between the tip and the surface. In the case of viscoelastic surfaces, the lateral deflection of a cantilever can be caused by (1) the shear friction between the tip and the surface (which may cause the plowing of the viscoelastic surface in front of the tip) and (2) the local tilt or slope of the surface. The nanoscale frictional properties of viscoelastic, compliant surfaces and their relations with the confined rheological properties have been investigated [12]. For this class of materials, the nonperturbative surface friction is a factor of the internal molecular relaxations near the surface, manifesting themselves in terms of the viscoelastic energy dissipative characteristics of the surface.
In another AFM mode, known as the force–volume mode, multiple force–distance curves are obtained across a surface. A force–distance curve (FDC) is a plot of net forces acting on an AFM tip as a function of the z-position of the piezo when the tip is brought into contact with the surface, slightly indents it, and withdraws from it [13,14]. By extracting nanomechanical properties such as the adhesion, the modulus, and the surface deformation from multiple FDCs, it will be possible to generate an image of the mechanical property of interest across a surface area.
The force tapping mode is a modified version of force–volume mode. The main difference between two operation modes is that, in force tapping mode, the z-piezo is driven by a sinusoidal wave, leading to the sinusoidal ramping (Figure 1a), while in force–volume mode, the z-piezo follows a triangular motion, leading to the triangular ramping [15]. In force tapping mode, during each period of probe modulation, a force–time plot is obtained by measuring the cantilever deflection as a function of time. A feedback loop system keeps the maximum applied force constant by adjusting the z-piezo extension (region C in Figure 1a,b). (The upper plot in Figure 1a shows one period of probe modulation in the force tapping mode.) In a real-time analysis, since the z-piezo position as a function of time is known, a FDC is generated from the force–time curve by replacing the time variable with the z-position of the modulation (Figure 1b).
The most prominent breakthrough of sinusoidal ramping is that the maximum indentation force can be applied and controlled more precisely, making it possible to study the nanomechanical properties of soft biomaterials [16]. This is mainly because the tip velocity decreases in a nonlinear fashion and eventually reaches zero at the maximum extension of the z-piezo. Therefore, more time will be given to the feedback loop system to adjust the z-piezo extension and control the maximum applied force at a higher resolution. Nevertheless, quantifying certain nanomechanical properties of soft biological surfaces, such as the elastic modulus, can still be challenging. Due to their natural surface roughness (with respect to the radius of curvature of a tip apex) and viscoelasticity, an accurate mapping of the elastic modulus of biological surfaces requires an implementation of more elaborate contact mechanics models.
An individual cotton fiber is an extremely elongated epidermal cell on the cottonseed’s surface, inside a cotton boll. A completely mature cotton fiber cell is made up of five distinguishable layers. The layers, from the innermost to the outermost, are the cell lumen, the secondary cell wall, the winding layer, the primary cell wall, and the cuticle [17]. The internal structure of a cotton fiber has been imaged with an AFM [18,19,20]. These studies have collected AFM images of a cotton fiber’s cross-section and surface—often scanned in liquid with the tapping mode—to investigate the dimension and the orientation of cellulose microfibrils deposited within the fiber’s secondary cell wall. The fiber’s outermost layer, the cuticle, is a few tens of nanometers-thick and is mostly composed of lipids, fatty acids, alcohols, and pectins (generally, hydrophobic compounds, collectively termed as cotton wax) [21,22]. The self-assembly and the growth of wax crystals, extracted from plant cell surfaces, have been imaged with an AFM [23]. These observations show that the morphology of wax crystals depends on the cell’s growth conditions, such as the temperature, the moisture content, the chemical composition of lipid compounds, and the structural template effects of the underlying substrate. The structure of individual macromolecules extracted from the cuticle layer of plant cells has been also studied with AFM nanotopography images. Round et al. used the AFM nanotopography images to investigate the branching, the molecular weight, and the molecular aggregation of individual pectin biopolymers extracted from the cuticle of a green tomato [24].
In addition to its biological functionalities, the cotton fiber’s cuticle has significant technological importance. Throughout various industrial fiber-processing stages, the waxy compounds function as a natural lubricant and facilitate the smooth motion of the fiber flow by reducing the inter-fiber friction and the friction between fiber and machine parts [11,25,26]. Studies have been conducted to investigate the nanoscale frictional and mechanical properties of the cuticular material of cotton fiber and other plant cells [27,28,29]. Round et al. isolated cutin biofilm from the cuticle of a tomato fruit and investigated its rheology with an AFM as a function of moisture content [28]. According to their results, the elastic modulus, obtained via the Hertz model, drastically decreased in a nonlinear fashion at the higher moisture content. Bhushan et al. examined the surface nanofriction of the plant leaves from four different species: lotus, colocasia, fagus, and magnolia [30]. Their results show that the hydrophilic leaves (fagus and magnolia) exhibited higher friction coefficients compared to that of the hydrophobic surfaces (e.g., lotus and colocasia). They explained that, for the hydrophilic surfaces, the contact area is expected to be larger, mainly due to the higher probability of water meniscus formation between the tip and the leaf surface [31].
Zhang et al. investigated the nanoscale frictional properties of cotton fibers [32]. The nanoscale frictional properties of synthetic and natural fibers can be used to evaluate the yield and the performance of ultrathin surface coatings. The authors showed that the nanoscale friction coefficient increased after multiple washes in the absence of fabric conditioner. They concluded that excessive washing cycles damage fibers’ surface and increase their surface roughness, leading to higher friction forces between the spherical AFM probe and the fibers’ surface. They also showed that the nanoscale friction experiments on cotton fiber conform to Amontons’ phenomenological law of friction F f = μ f L , where, Ff is the friction force, L is the applied normal force, and μ f is the friction coefficient.
The main objective of this study was to quantify and compare the surface nanomechanical and nanoscale frictional properties of two different samples of cotton fiber which are known to have statistically distinct macroscale frictional properties. For this reason, the fiber samples were characterized with an AFM, both in contact and force tapping operation modes. Fundamental single-asperity nanoscale friction experiments on cotton fibers’ surface can be used to better understand the linkage between the nanoscale and macroscale frictional properties of fibers.
This study serves as a preliminary investigation into the surface characterization of cotton fibers with an AFM which opens new research avenues for further investigations on this field.

2. Materials and Methods

2.1. Materials

Two different varieties of cotton fiber were selected from a larger set of samples for a comparative study: ST 4747GLB2 (Bayer Crop Science, Monheim am Rhein, Germany) and BRS 293 (experimental strain). The samples were labeled A and B, respectively. In the previous study by the authors, these two varieties showed statistically distinct macroscale frictional properties [10]. Cotton plants used to produce fiber for this research were grown under the same conditions [10]. After fiber samples were collected, their dimensional and mechanical properties were measured using the standard procedures for cotton fiber classification and testing. The physical characteristics of the fiber samples are summarized in Table 1. The macroscale friction coefficients of the fibers were measured using a sliding friction tester in which an assembly of randomly oriented fibers was pulled against a nominally flat metallic plate under varying normal force L [10]. The friction coefficient μ f was calculated by fitting the data to the F f = μ f L model. Micronaire is a measure of the air permeability through a known mass of compressed assembly of fiber. The micronaire value is highly correlated to the interaction of a fiber’s linear density and maturity. Strength, in g/tex, is the normalized force required to break a bundle of fiber (tex is the weight in g of 1000 m of fiber). The +b is comparable to the b* dimension in the CIE color space, which determines the degree of yellowness of the bulk of fibers. The micronaire, length, strength, elongation, and +b of the fiber varieties were measured with Uster®’s High Volume Instrument. Eight individual fibers were randomly selected from each variety for AFM experiments.

2.2. Sample Preparation

Individual fibers were mounted on microscope glass slides. First, one end of a fiber was fixed to a glass slide using a thin layer of glue. After the glue dried, the fiber was gently stretched to remove its natural crimp. While the fiber was extended, multiple droplets of glue were applied along its length. The distance between the two fixed points along the fiber length was 1 ± 0.5 mm. With this technique, the fiber was firmly immobilized on the substrate, prohibiting any significant drift or creep during AFM imaging. In order to avoid any contact between a probe chip and the glue joints, we made certain that the height of the joints did not exceed the height of the fiber, which is in the order of ~4 µm.

2.3. Nanomechanical Properties Measurements

Nanotopography and nanomechanical property images of the fibers’ surface were obtained using a multimode AFM (Bruker’s Dimension Icon, Billerica, MA, USA) in PeakForce quantitative nanomechanical property mapping (PF-QNM) mode. The nanomechanical experiments were performed using a blunt probe (Team Nanotec GmbH, Villingen-Schwenningen, Germany) with a radius of curvature R = 40 nm and a nominal spring constant C = 0.06 N/m. The cone angle of the tip was 40° and its height was 9 µm. Before each set of experiments, the laser beam was aligned on the back of the cantilever. The photodetector position was adjusted accordingly to maximize the gain signal. The deflection sensitivity of the cantilevers was calibrated by making five FDCs on a sapphire surface. The cantilever spring constant was determined by recording the thermal noise power spectrum of the cantilevers. The approximate regions on the fibers’ surface for the purpose of force measurement experiments were determined with the built-in video camera (508–4010× magnification range and a corresponding field of view of 190–1500 µm). The experiments were conducted in a 5 µm × 5 µm scan size with 512 lines per scan. Although a higher ramping resolution could be achieved in force tapping mode, but by maintaining the ramping frequency at the recommended level, we made certain that mechanical property measurements were not affected by time-dependent factors such as the increase in the localized stiffness due to constrained molecular relaxations. All measurements were carried out in the air. Three locations per fiber were imaged. The scan rate was adjusted to 0.5 Hz (= tip velocity of 5 µm/s). To minimize tip contamination and wear, the PeakForce setpoint was fixed to 10 nN. In order to reduce the contact time between the tip and sample, the PeakForce amplitude was set to 300 nm. The z-piezo was set to modulate at 1 kHz, which is the recommended value for viscoelastic surfaces. The produced topography (height sensor), PeakForce error, adhesion, and deformation images were collected for further statistical analyses.

2.4. Nanoscale Friction Measurements

The surface topography and friction images of the fiber samples were obtained using a multimode AFM (Bruker’s Dimension Icon) in contact mode. The probe tip used for the friction experiments was the same tip used in the force measurement experiments (R = 40 nm and C = 0.06 N/m). Three locations per fiber were scanned with a scan size of 2 µm × 2 µm. The scan speed, in the fast scan direction (perpendicular to the primary cantilever axis), was adjusted to 4 µm/s (calculated from 2 × scan length × scan frequency). Similar to nanomechanical measurements, the scanning speed was maintained at the recommended level, making certain that the friction measurements were not affected by uncontrolled factors such as the friction-derived heating and the increase in the localized stiffness. The friction signal, in volts, was measured by taking the average difference between the lateral deflection of the cantilever obtained from right-to-left and left-to-right scanning. In this comparative study, in which all the measurements were obtained with the same probe, the conversion of the lateral deflection of the cantilever to the absolute frictional force was not necessary. Contact forces were varied by changing the vertical deflection of the cantilever during scanning. The friction measurements were carried out in the air, under the normal force of 10 to 100 nN.

2.5. Estimating the Contact Area

According to the Jonson–Kendall–Roberts (JKR) model, the tensile force (a.k.a., pull-off force) required to separate surfaces from contact can be obtained from
F a d = 3 / 2 π R W ,
where F a d is adhesive force, R is the radius of curvature of the tip, and W is the adhesion energy. For the elastic contact of two different solids, the adhesion energy can be estimated from W = γ A + γ B + γ A B , where γ denotes the interfacial energies between the materials A, B, and air. Solving Equation (1) for W results:
W = 2 F a d / 3 π R .
Several studies have successfully employed the JKR model to estimate the adhesion energy of a wide range of viscoelastic polymers [33,34,35,36]. The adhesion energy obtained from the JKR model on polymer solids correlates well with those obtained from contact angle methods. In this study, we used Equation (2) to determine the adhesion energy of the fiber samples. Afterwards, different contact mechanics theories were implemented to estimate the contact area between the AFM tip and the surface of the fiber. Contact mechanics is defined as the study of the mechanics of solids spreading their interfacial contact area under normal force [14]. According to Hertz contact theory, for a sphere of radius R 1 in contact with a flat surface ( R 2 = ), the true contact area is in the shape of a circle. Under elastic deformation, the contact radius a can be predicted as
a = R 1 L E t 3 ,
where E t is the reduced Young’s modulus. The reduced Young’s modulus can be obtained from
1 E t = 3 4 1 v s 2 E s + 1 v a 2 E a ,
where ν a , E a , ν s , and E s are the Poisson’s ratios and Young’s moduli of the asperity and the surface, respectively. In the Hertz theory, the adhesion between the two surfaces is neglected. Jonson et al. and Derjaguin et al. modified the Hertz model by taking into account the attractive forces between surfaces. In the JKR model, the contact radius a can be obtained from the following equation:
a 3 = R 1 E t 3 2 π R 1 W + L + 3 2 π R 1 W 2 .
The JKR model is mostly applicable for soft materials with large adhesion. The Derjaguin–Muller–Toporov (DMT) model, on the other hand, is applied in the case of stiff samples with small adhesion. The contact radius in the DMT model can be obtained from the following equation:
a 3 = R 1 E t L + 2 π R 1 W .
Equations (3), (5) and (6) were used to estimate the real contact area A r e a l = π a 2 of the fibers as a function of L and R.

3. Results and Discussion

The AFM limits in producing height images of a cotton fiber’s surface at different length scales are presented in Figure 2a. The absence of noise in the images shows that the described methodology successfully immobilized the fibers during scanning experiments. For more clarification, the main differences in the working principles of contact mode and force tapping mode (a.k.a., Bruker’s PeakForce QNM) are illustrated in Figure 2b. In contact mode, the tip stays in contact with the surface throughout an entire raster scanning process, producing continuous scan lines. In the force tapping mode, however, each scan line is regenerated by connecting multiple points along the scanning direction of the probe. Each point in the three-dimensional space is the result of one tip–surface interaction during one sinusoidal modulation of the probe. The distance between the points determines the image resolution. The corresponding FDC at each interaction (Figure 2c) can be recorded for further mechanical property analyses.
Figure 3a shows a scanning electron microscope image of two typical cotton fibers. Typical, false-color images of the nanomechanical properties of the fibers’ surface—produced in force tapping mode—are presented in Figure 3b.
The 5 μm × 5 μm surface topography of the fibers’ surface obtained in force tapping mode are shown in Figure 4 (four scans per each sample are presented). The scanned areas could have been acquired from different regions of the fibers’ surface. Since a mature cotton fiber is a convoluted structure with a bean-shape cross-section (Figure 2a and Figure 3a), the scanned areas can be concave, convex, or relatively flat. A sequence of furrows and ridges is evident on the fibers’ surface [19].
Figure 5 shows the 2 μm × 2 μm surface topography images of the fibers’ surface obtained in contact mode (the corresponding friction images of the scanned areas, under different normal forces, are shown in Figure 6 and Figure 7). In addition to the furrows and ridges on the fibers’ surface, a series of irregular-shape, plaque-like surface aggregates is evident at this length scale [19]. The granular surface aggregates are more abundant on sample B fibers (an average of 15 counts per scan area for sample B compared to an average of 1 count per scan area for sample A). While their exact nature is unknown, they can be due to the self-assembly and growth of wax crystals on the cuticle surface.
The friction images of the fibers obtained under 10 nN normal force (with respect to the 40-nm radius of curvature of the AFM tip apex) are shown in Figure 6. Under this normal force, sample B shows higher nanoscale friction on average. The mean friction signal for sample B is 0.50 ± 0.01 V compared to the lower value of 0.33 ± 0.01 V for sample A. Figure 7 shows the friction images of the fibers obtained under the higher normal force of 50 nN; as can be noted, the fibers’ surface from sample A exhibits significantly higher nanoscale friction compered to sample B.
The local variations in the nanoscale friction signal, which are more evident on sample B fibers, are due to topographical effects and can be explained by the ratchet mechanism [37]. As the tip encounters an asperity with a given slope during raster scanning, it climbs against it and transmits a larger lateral deflection signal to the position-sensitive detector. However, the lateral deflection signal is in part due to the variations in the slope of the surface. The local variations are often referred to as the ratchet-mechanism component of friction. Even though the trace and retrace friction signals have been subtracted, the local variations are still evident in the produced images.
Figure 8 shows the histograms of the friction experiments under 50 nN normal force and their corresponding Gaussian fit. The distributions of the two samples are quite distinguishable. The average friction force is 1.48 ± 0.02 V for sample A, compared to the significantly lower value of 0.80 ± 0.02 V for sample B fiber.
The relationship between the nanoscale friction and normal force for tested fibers is depicted in Figure 9. With an increase in normal force, the friction on sample B increases to a certain extent and then reaches a plateau. The sublinear relationship between the nanoscale friction and the normal force in a nonperturbative state is expected. In the absence of any surface wear, the friction is linearly proportional to the real contact area, which itself holds a sublinear relationship with the applied normal force [38]. Therefore, a model of the form y = b x n (where, b and n are the hyperparameters and 0 < n 1 is the inequality constraint) was fitted to the friction data using the method of least squares in MATLAB Optimization Toolbox™ (MathWorks, Natick, MA, USA).
The theoretical basis behind the fitted model is as follows. It is known that, according to the adhesion theory of friction,
F f = τ ¯ A r e a l ,
where τ ¯ is the effective shear strength of contacting bodies. Under nonperturbative friction, and if we assume that the contacting bodies deform in an elastic Hertzian regime, the real contact area can be obtained as
A r e a l = π R 1 L E t o t 3 2 .
By substituting A r e a l into Equation (7), one can obtain
F f = τ ¯ π R 1 L E t o t 2 3 .
Equation (9) can then be generalized to the following form:
F f = b L n ,
where b = τ ¯ π R 1 / E t o t 2 3 is a constant related to the shape, the mechanical properties, and the shear strength of the contact bodies.
By fitting Equation (10) to the friction data for sample B, b ¯ = 0.51 ± 0.05 and n ¯ = 0.25 ± 0.04 are obtained. The average of the coefficient of determination, r ¯ 2 , of the fitted models is 0.99, which is very close to the perfect r-squared value of 1. The extremely small average p-value (<0.0001) for the model hyperparameters indicates that both statistics are statistically significant at a 99% level of confidence (i.e., the model parameters are not equal to zero at a very high level of certainty).
The friction plot of sample A, however, does not reach a plateau within the range of applied normal force. Under the low normal forces, the friction increases in a nonlinear fashion but at a smaller rate compared to sample B ( b ¯ , n ¯ , and r ¯ 2 values of 0.30, 0.45, and 1, respectively). With further increase in the normal force, the friction shifts to the higher values and exceeds the mean frictions for sample B. Under these normal forces (i.e., 30 nN and higher), the linear regression model is a perfect fit to the data with r ¯ 2 = 0.99 .
The shift, which can be interpreted as a transition phase in the friction profile of a surface, is due to the occurrence of surface damage and wear debris formation. It has been reported that, upon the occurrence of wear or surface debris, the friction force deviates from the nonlinearity regime and follows a simple linear regression model [39]. With the development of plastic deformation during a sliding friction process, two situations are likely: either the wear debris remain between the contacting bodies and reduce the overall friction via rolling friction mechanism, or the displaced molecules pile up on the sides of the contacting bodies and cause higher friction by increasing the real contact area. Direct observation of either situation is not feasible with current commercial AFM technologies. To the authors’ knowledge, only Surface Apparatus can provide the real-time projection of nanoscale deformations between surfaces during friction experiments of thin films. In either situation, 0 < n < 1 , as presented in Equation (10), reaches unity, and Equation (7) takes the form of Amontons’ law of friction, F f = μ f L , stating that the friction force is linearly proportional to the normal force.
In AFM experiments, wear debris formed after friction experiments can be analyzed by rescanning the wider area of the previously tested region. Figure 10a shows a typical molecular pile-up with the height of 140 nm on the sides of the previously scanned area from sample A. The pile-up formation, which is not evident on sample B surfaces (Figure 10b), further confirms that the transition phase due to molecular displacement has occurred on sample A surfaces, evidently between 20 and 30 nN normal forces.
These results suggest that when the friction profile data conform to the Hertzian component of friction, those experiments might have been performed in a nonperturbative state, on the same material (i.e., lipid compounds on fiber’s surface as depicted in Figure 10c). However, once the friction curve deviates from sublinearity, the remaining experiments might have been performed beyond elastic limits of the thin cuticle layer, probably on the layer underneath it (i.e., the primary cell wall as depicted in Figure 10d). Based on these conclusions, we hypothesize that fibers from sample B with lower macroscale friction coefficients might have been coated with a thicker cuticle layer since we did not observe the transition phase in their friction profile, nor the trace of plastic deformation on their wear images.
Figure 11 shows the adhesion images of the fibers’ surface, obtained in PeakForce QNM® mode. As can be noted, the average adhesion is significantly higher for sample B. The mean tip–surface adhesive force for the sample B fibers is 22.8 ± 12.0 nN compared to the lower value of 15.1 ± 7.0 nN for the sample A. The p-value of zero indicates that the difference between the adhesion means is statistically significant at the 5% significance level.
The positive correlation between the adhesion (performed under the low normal force of 10 nN) and the friction under the same low normal force is in agreement with the adhesion theory of friction, which states that, during a nonperturbative single-asperity friction process, the friction force is adhesion-controlled ( F f = τ ¯ A r e a l ). In other words, the greater the single-asperity adhesion force, the larger the shear strength of the contacting asperities will be.
The magnitude of the adhesive force between an AFM tip and a surface depends chiefly on the likelihood of water meniscus formation (i.e., the capillary forces), the chemistry of surfaces, and the real contact area between contacting asperities. In the case of cotton fiber’s surface, the cuticle layer is essentially a hydrophobic compound. This makes the water meniscus formation hypothesis improbable. Therefore, we hypothesize that it is the chemistry of the cuticle layer (such as the hydrocarbon chain length of the lipids) which determines the overall attractive forces, the surface energies, and the mechanical properties of the fibers’ surface (i.e., its stiffness, Poisson ratio, and internal friction). It remains to be studied whether there is any direct correlation between the adhesion force and the thickness of the cuticle layer.
In addition to the average adhesive force for each scanned area, the local variation in the adhesive force can be noted from the adhesion images (Figure 11a). The local variations tend to resemble the pattern of furrows and ridges on the fibers’ surface, mainly due to the folds on the primary cell wall (Figure 3b). In general, if other factors are held constant, the adhesive force is larger for concave curvatures. This is because of their larger real contact area, which increases the probability of molecular adhesive bond formation (mostly in the form of van der Waals forces) between the tip and the surface.
Figure 12 shows the maps of the surface deformation of the fiber samples, obtained in PeakForce QNM® mode, and distributions of surface deformation for all other tested fibers. The mean maximum penetration depth of the tip is considerably higher for sample A fibers. The mean maximum deformation for the sample A fibers is 20.9 ± 4.7 nm, compared to the lower value of 3.2 ± 1.4 nm for the sample B fibers. It has to be noted that the nanoscale deformation can be broken down into multiple physical components. According to JKR theory, under the elastic regime, the indentation depth d is related to R, a, the Young’s modulus of the surface E S , and W:
d = a 2 R 2 3 6 π W a E t .
Among these variables, R was constant throughout the experiments. However, a, W, and E S could vary in each experiment. The contact radius a depends on R, E S , L, and W. Since the applied normal force was kept constant, the variation in the surface deformation d could only be explained in terms of W and E S ; that is, the increase in adhesion energy and/or reduction in the local stiffness increase the surface deformation. Thus, the higher surface deformation of sample A fibers is attributed to the combination of both its surface adhesion and stiffness. Both of these surface attributes are related to the chemistry and composition of the cuticle layer as functions of temperature and moisture content.
Since the real contact area between a tip and a surface cannot be measured in an AFM experiment, the best approach to estimate the real contact area is to employ different contact mechanics theories such as the Hertz, the JKR, and the DMT. The input average adhesion energy for sample A and B fibers (as computed from Equation (2)) were 0.08 and 0.12 N/m, respectively. Figure 13 shows the estimated real contact area between a silicon nitride tip with a radius of curvature of 40 nm and a lipid surface as a function of the normal force and the adhesion energy. Each line in the plot shows the variation in the theoretical real contact area as a function of normal force for a given adhesion energy. The adhesion energy varies from 0 to 30 N/m. The variation in the real contact area between two samples was not noticeable at the selected length scale; therefore, a single line of W = 0.1 is emphasized to represents both samples.
For instance, under the normal force of 50 nN, based on the JKR theory, the estimated contact radius between a lipid layer and a silicon nitride tip with a radius of curvature of 40 nm is ~110 nm. Under the same conditions, the estimated contact radius based on the DMT theory is only ~6.5 nm. Among different contact theories, the JKR always overestimates the real contact area. This is because the JKR takes into account the short-range adhering force between the contacting surfaces once the contact is made, while the DMT only considers the long-range attractive forces before the contact is made. The Hertzian model underestimates the real contact area since it ignores attractive forces between the surfaces. It can be noted that all these models can be fitted to adhesion-controlled, nonperturbative, single-asperity friction data since their shape resembles friction curves very closely. In this study, we fitted the Hertz model to the fibers’ surface friction data mainly due to its simplicity. Among these theories, it is reported that the JKR model provided the most accurate estimate for the real contact area of biological and soft surfaces [40].

4. Conclusions

To better understand the origin of the frictional properties of cotton fiber assemblies, surface nanotribological properties of two cotton fiber samples were investigated at the nanoscale with an AFM. The surface topography of the fibers was analyzed in both contact and force tapping modes. The fibers’ surface from both samples are characterized by series of furrows and ridges, while the fibers’ surface from sample B are also characterized by certain granular-like surface aggregates.
The surface nanotribological properties of the fiber samples were examined under various normal forces. Friction data from both samples were fitted to the equation of single-asperity, adhesion-controlled friction ( F f = b L n ) with high accuracy. The friction curves of sample A, however, show a characteristic transition phase during which the adhesion-controlled fit transforms into Amontons’ law of friction. The high r ¯ 2 values, along with the observed wear marks in height images of sample A, confirm that the transition phases are due to the displacement of the cuticle layer molecules. These observations lead us to hypothesize that fibers from sample B are naturally coated with a thicker layer of the cuticle. The hypothesis for the thickness of the cuticle layer of the samples may partially explain the lower macroscale friction of sample B fibers by the fact that a thicker cuticular membrane provides more lubrication effects.
The nanoscale adhesion and deformation of the fibers were also investigated in force tapping mode under the low normal force of 10 nN (with respect to the 40-nm radius of curvature of the AFM tip apex). The sample B fibers show higher average adhesion and much lower nanoscale deformation. With reference to the adhesion theory of the friction ( F f = τ ¯ A r e a l ), the higher average adhesion for sample B fibers is consistent with its higher nanoscale friction under the low normal force of 10 to 20 nN. We hypothesize that the differences in the average nanomechanical properties of two samples are related to the chemistry and the mechanical properties of their cuticle layer. The local variation in the adhesion and the deformation images are due to the shape of folds and asperities on the fibers’ surface.
Lastly, by computing the average adhesion energy for both samples, the theoretical real contact area between the probe and the fibers as a function of normal force was estimated. It is explained that, with correct assumptions of the adhesion energy, the mechanical properties of contact bodies, and their shape, the contact mechanics models can be precisely fitted to an adhesion-controlled, nonperturbative friction process.

Author Contributions

F.H. conducted the experiments, analyzed the data, and wrote the manuscript. J.A.T. gathered the funding, supervised the project, and guided the research.

Funding

This work was supported in part by the Cotton Incorporated under Grant 05-755. This work was supported in part by John Alexander Thomasson of the Department of Biological and Agricultural Engineering at Texas A&M University.

Acknowledgments

The authors would like to acknowledge Cotton Incorporated for its financial support throughout the project. We would also like to thank the staff at Material Characterization Facility (MCF) of Texas A&M University for giving us access to the AFM and its software. Special thanks goes to Wilson Serem of Texas A&M University for providing AFM training.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Perry, S.S. Scanning probe microscopy measurements of friction. MRS Bull. 2004, 29, 478–483. [Google Scholar] [CrossRef]
  2. Hosseinali, F.; Thomasson, J. Multiscale frictional properties of cotton fibers: A review. Fibers 2018, 6, 49. [Google Scholar] [CrossRef]
  3. Bargel, H.; Koch, K.; Cerman, Z.; Neinhuis, C. Structure–function relationships of the plant cuticle and cuticular waxes—A smart material? Funct. Plant Biol. 2006, 33, 893–910. [Google Scholar] [CrossRef]
  4. Domínguez, E.; Cuartero, J.; Heredia, A. An overview on plant cuticle biomechanics. Plant Sci. 2011, 181, 77–84. [Google Scholar] [PubMed]
  5. Domínguez, E.; Heredia-Guerrero, J.A.; Heredia, A. The biophysical design of plant cuticles: An overview. New Phytol. 2011, 189, 938–949. [Google Scholar] [CrossRef] [PubMed]
  6. Mirabet, V.; Das, P.; Boudaoud, A.; Hamant, O. The role of mechanical forces in plant morphogenesis. Annu. Rev. Plant Biol. 2011, 62, 365–385. [Google Scholar] [CrossRef] [PubMed]
  7. Milani, P.; Braybrook, S.A.; Boudaoud, A. Shrinking the hammer: Micromechanical approaches to morphogenesis. J. Exp. Bot. 2013, 64, 4651–4662. [Google Scholar] [CrossRef]
  8. Routier-Kierzkowska, A.L.; Weber, A.; Kochova, P.; Felekis, D.; Nelson, B.J.; Kuhlemeier, C.; Smith, R.S. Cellular force microscopy for in vivo measurements of plant tissue mechanics. Plant Physiol. 2012, 158, 1514–1522. [Google Scholar] [CrossRef]
  9. Li, Q.; Li, Y.; Zhu, L.; Xing, B.; Chen, B. Dependence of plant uptake and diffusion of polycyclic aromatic hydrocarbons on the leaf surface morphology and micro-structures of cuticular waxes. Sci. Rep. 2017, 7, 46235. [Google Scholar]
  10. Hosseinali, F.; Thomasson, J.A. Variability of fiber friction among cotton varieties: Influence of salient fiber physical metrics. Tribol. Int. 2018, 127, 433–445. [Google Scholar] [CrossRef]
  11. Sui, R.; Thomasson, J.A.; Byler, R.K.; Boykin, J.C.; Barnes, E.M. Effect of machine-fiber interaction on cotton fiber quality and foreign-matter particle attachment to fiber. J. Cotton Sci. 2010, 153, 145–153. [Google Scholar]
  12. Hammerschmidt, J.A.; Moasser, B.; Gladfelter, W.L.; Haugstad, G.; Jones, R.R. Polymer viscoelastic properties measured by friction force microscopy. Macromolecules 1996, 29, 8996–8998. [Google Scholar] [CrossRef]
  13. Burnham, N.A.; Colton, R.J.; Pollock, H.M. Interpretation of force curves in force microscopy. Nanotechnology 1993, 4, 64–80. [Google Scholar] [CrossRef]
  14. Butt, H.J.; Cappella, B.; Kappl, M. Force measurements with the atomic force microscope: Technique, interpretation and applications. Surf. Sci. Rep. 2005, 59, 1–152. [Google Scholar] [CrossRef] [Green Version]
  15. Dufrêne, Y.F.; Martínez-Martín, D.; Medalsy, I.; Alsteens, D.; Müller, D.J. Multiparametric imaging of biological systems by force-distance curve–based AFM. Nat. Methods 2013, 10, 847–854. [Google Scholar] [CrossRef]
  16. Schillers, H.; Medalsy, I.; Hu, S.; Slade, A.L.; Shaw, J.E. PeakForce Tapping resolves individual microvilli on living cells. J. Mol. Recognit. 2016, 29, 95–101. [Google Scholar] [CrossRef]
  17. Smith, C.W.; Cothren, J.T. Cotton: Origin, History, Technology, and Production; John Wiley and Sons: Hoboken, NJ, USA, 1999; ISBN 0471180459. [Google Scholar]
  18. Maxwell, J.M.; Gordon, S.G.; Huson, M.G. Internal structure of mature and immature cotton fibers revealed by scanning probe microscopy. Text. Res. J. 2003, 73, 1005–1012. [Google Scholar] [CrossRef]
  19. Pesacreta, T.C.; Carlson, L.C.; Triplett, B.A. Atomic force microscopy of cotton fiber cell wall surfaces in air and water: Quantitative and qualitative aspects. Planta 1997, 202, 435–442. [Google Scholar] [CrossRef]
  20. Lee, I.; Evans, B.R.; Woodward, J. The mechanism of cellulase action on cotton fibers: Evidence from atomic force microscopy. Ultramicroscopy 2000, 82, 213–221. [Google Scholar] [CrossRef]
  21. Vaughn, K.C.; Turley, R.B. The primary walls of cotton fibers contain an ensheathing pectin layer. Protoplasma 1999, 209, 226–237. [Google Scholar] [CrossRef]
  22. Hartzell-Lawson, M.M.; Hsieh, Y.L. Characterizing the noncellulosics in developing cotton fibers. Text. Res. J. 2000, 70, 810–819. [Google Scholar] [CrossRef]
  23. Koch, K.; Neinhuis, C.; Ensikat, H.; Barthlott, W. Self assembly of epicuticular waxes on living plant surfaces imaged by atomic force microscopy (AFM). J. Exp. Bot. 2004, 55, 711–718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Round, A.N.; MacDougall, A.J.; Ring, S.G.; Morris, V.J. Unexpected branching in pectin observed by atomic force microscopy. Carbohydr. Res. 1997, 303, 251–253. [Google Scholar] [CrossRef]
  25. El Mogahzy, Y.; Broughton, R.; Hong Guo, H.; Taylor, R.A. Evaluating staple fiber processing propensity. Text. Res. J. 1998, 68, 835–840. [Google Scholar] [CrossRef]
  26. Sui, R.; Thomasson, J.A.; Ge, Y.; Morgan, C. Multispectral sensor for in-situ cotton fiber quality measurement. Trans. ASABE 2008, 51, 2201–2208. [Google Scholar] [CrossRef]
  27. Chatterjee, S.; Sarkar, S.; Oktawiec, J.; Mao, Z.; Niitsoo, O.; Stark, R.E. Isolation and biophysical study of fruit cuticles. J. Vis. Exp. 2012, 61, e3529. [Google Scholar] [CrossRef] [PubMed]
  28. Round, A.N.; Yan, B.; Dang, S.; Estephan, R.; Stark, R.E.; Batteas, J.D. The Influence of water on the nanomechanical behavior of the plant biopolyester cutin as studied by AFM and solid-state NMR. Biophys. J. 2000, 79, 2761–2767. [Google Scholar] [CrossRef]
  29. Batteas, J.D.; Stark, R.E. 16—Surface and interfacial studies of plant biopolymers. In Molecular Interfacial Phenomena of Polymers and Biopolymers; Woodhead Publishing: Saxton, UK, 2005; pp. 580–608. ISBN 9781855739284. [Google Scholar]
  30. Bhushan, B.; Jung, Y.C. Micro-and nanoscale characterization of hydrophobic and hydrophilic leaf surfaces. Nanotechnology 2006, 17, 2758–2772. [Google Scholar] [CrossRef]
  31. Weeks, B.L.; Vaughn, M.W.; DeYoreo, J.J. Direct imaging of meniscus formation in atomic force microscopy using environmental scanning electron microscopy. Langmuir 2005, 21, 8096–8098. [Google Scholar] [CrossRef]
  32. Zhang, Z.; Fletcher, I.W.; Hurley, C.R.; Boardman, C.; Doyle, P.; Leggett, G.J.; Poutch, F.; Turner, P.S.; Watson, G.S. Morphological and quantitative frictional measurements of cotton fibres using friction force microscopy. J. Mater. Chem. 2010, 20, 8531–8538. [Google Scholar] [CrossRef]
  33. Merrill, W.W.; Pocius, A.V.; Thakker, B.V.; Tirrell, M. Direct measurement of molecular level adhesion forces between biaxially oriented solid polymer films. Langmuir 1991, 7, 1975–1980. [Google Scholar] [CrossRef]
  34. Mangipudi, V.; Tirrell, M.; Pocius, A.V. Direct measurement of the surface energy of corona-treated polyethylene using the surface forces apparatus. Langmuir 1995, 11, 19–23. [Google Scholar] [CrossRef]
  35. Mangipudi, V.; Tirrell, M.; Pcius, A.V. Direct measurement of molecular level adhesion between poly(ethylene terephthalate) and polyethylene films: Determination of surface and interfacial energies. J. Adhes. Sci. Technol. 1994, 8, 1251–1270. [Google Scholar] [CrossRef]
  36. Tirrell, M. Measurement of interfacial energy at solid polymer surfaces. Langmuir 1996, 12, 4548–4551. [Google Scholar] [CrossRef]
  37. Bhushan, B. Nanotribology and nanomechanics. Wear 2005, 259, 1507–1531. [Google Scholar] [CrossRef]
  38. Urbakh, M.; Klafter, J.; Gourdon, D.; Israelachvili, J. The nonlinear nature of friction. Nature 2004, 430, 525–528. [Google Scholar] [CrossRef]
  39. Homola, A.M.; Israelachvili, J.N.; McGuiggan, P.M.; Gee, M.L. Fundamental experimental studies in tribology: The transition from “interfacial” friction of undamaged molecularly smooth surfaces to “normal” friction with wear. Wear 1990, 136, 65–83. [Google Scholar] [CrossRef]
  40. Tsukruk, V.V.; Huang, Z.; Chizhik, S.A.; Gorbunov, V.V. Probing of Micromechanical Properties of Compliant Polymeric Materials. J. Mater. Sci. 1998, 33, 4905–4909. [Google Scholar] [CrossRef]
Figure 1. (a) Position of z-piezo (upper plot) and vertical deflection of cantilever (lower plot) as functions of time during one modulation of tip in force-tapping mode; (b) vertical deflection of the tip as a function of z-piezo displacement: A—probe tip above sample surface, B—tip snaps into contact with surface, C—ultimate contact is made and tip indents surface at a predefined PeakForce™ setpoint, D—tip–surface separation (adhesion force), E—z-piezo contracts and probe springs back to starting point; (c) atomic force microscope (AFM) probe in contact with the cotton fiber’s surface (view from the AFM’s built-in video camera).
Figure 1. (a) Position of z-piezo (upper plot) and vertical deflection of cantilever (lower plot) as functions of time during one modulation of tip in force-tapping mode; (b) vertical deflection of the tip as a function of z-piezo displacement: A—probe tip above sample surface, B—tip snaps into contact with surface, C—ultimate contact is made and tip indents surface at a predefined PeakForce™ setpoint, D—tip–surface separation (adhesion force), E—z-piezo contracts and probe springs back to starting point; (c) atomic force microscope (AFM) probe in contact with the cotton fiber’s surface (view from the AFM’s built-in video camera).
Fibers 07 00064 g001
Figure 2. (a) Surface topography images of a cotton fiber obtained at different scan sizes; as can be seen, the cotton fiber has a bean-shape cross-section. (b) Different scan lines produced from different AFM modes. (c) An array of force–distance curves (FDCs) produced along a scan line in PeakForce tapping mode.
Figure 2. (a) Surface topography images of a cotton fiber obtained at different scan sizes; as can be seen, the cotton fiber has a bean-shape cross-section. (b) Different scan lines produced from different AFM modes. (c) An array of force–distance curves (FDCs) produced along a scan line in PeakForce tapping mode.
Fibers 07 00064 g002
Figure 3. (a) Scanning electron microcopy image of two typical cotton fibers; (b) Typical, 5 μm × 5 μm, nanotopographical and nanomechanical property images of a cotton fiber’s surface obtained in force tapping mode.
Figure 3. (a) Scanning electron microcopy image of two typical cotton fibers; (b) Typical, 5 μm × 5 μm, nanotopographical and nanomechanical property images of a cotton fiber’s surface obtained in force tapping mode.
Fibers 07 00064 g003
Figure 4. The 5 μm × 5 μm three-dimensional surface topography of sample A (upper panel) (a) and sample B (lower panel) (b) fibers obtained in force tapping mode.
Figure 4. The 5 μm × 5 μm three-dimensional surface topography of sample A (upper panel) (a) and sample B (lower panel) (b) fibers obtained in force tapping mode.
Fibers 07 00064 g004
Figure 5. The 2 μm × 2 μm three-dimensional surface topography of sample A (upper panel) (a) and sample B (lower panel) (b) fibers obtained in contact mode; the granular surface aggregates can be due to the self-assembly and growth of wax crystals.
Figure 5. The 2 μm × 2 μm three-dimensional surface topography of sample A (upper panel) (a) and sample B (lower panel) (b) fibers obtained in contact mode; the granular surface aggregates can be due to the self-assembly and growth of wax crystals.
Fibers 07 00064 g005
Figure 6. Friction images of sample A (upper panel) (a) and sample B (lower panel) (b) fibers obtained under 10 nN normal force; surface of fibers from sample B exhibited significantly higher nanoscale friction.
Figure 6. Friction images of sample A (upper panel) (a) and sample B (lower panel) (b) fibers obtained under 10 nN normal force; surface of fibers from sample B exhibited significantly higher nanoscale friction.
Fibers 07 00064 g006
Figure 7. Friction images of sample A (upper panel) (a) and sample B (lower panel) (b) fibers obtained under the normal force of 50 nN; fibers’ surface from sample A showed significantly higher nanoscale friction.
Figure 7. Friction images of sample A (upper panel) (a) and sample B (lower panel) (b) fibers obtained under the normal force of 50 nN; fibers’ surface from sample A showed significantly higher nanoscale friction.
Fibers 07 00064 g007
Figure 8. Friction force histograms and their corresponding Gaussian fit for the friction experiments under 50 nN normal force; the friction force distributions of the two samples are well-separated.
Figure 8. Friction force histograms and their corresponding Gaussian fit for the friction experiments under 50 nN normal force; the friction force distributions of the two samples are well-separated.
Fibers 07 00064 g008
Figure 9. Relationship between single-asperity friction and normal force for tested surfaces; the transition phase from F f = b L n to F f = μ f L for sample A denotes the state at which the cuticle layer molecules are displaced from the fibers’ surface.
Figure 9. Relationship between single-asperity friction and normal force for tested surfaces; the transition phase from F f = b L n to F f = μ f L for sample A denotes the state at which the cuticle layer molecules are displaced from the fibers’ surface.
Fibers 07 00064 g009
Figure 10. (a) Typical wear mark after the set of friction experiments on sample A; the plot shows the height profile along the dashed line. (b) Typical surface profile of sample B after the fibers for the set of friction experiments; no evidence of wear marks can be detected on the image. (c) Hypothetical schematic of the tip–surface interaction for fibers with a relatively thick cuticular layer under high normal forces. (d) Hypothetical schematic of the tip–surface interaction for fibers with a thin cuticular membrane under high normal forces.
Figure 10. (a) Typical wear mark after the set of friction experiments on sample A; the plot shows the height profile along the dashed line. (b) Typical surface profile of sample B after the fibers for the set of friction experiments; no evidence of wear marks can be detected on the image. (c) Hypothetical schematic of the tip–surface interaction for fibers with a relatively thick cuticular layer under high normal forces. (d) Hypothetical schematic of the tip–surface interaction for fibers with a thin cuticular membrane under high normal forces.
Fibers 07 00064 g010
Figure 11. (a) Adhesion images of the fiber samples obtained in the PeakForce tapping mode; (b) the adhesive force distributions and their Gaussian fit. The arithmetic means for both varieties, the significance level, and the p-value are presented.
Figure 11. (a) Adhesion images of the fiber samples obtained in the PeakForce tapping mode; (b) the adhesive force distributions and their Gaussian fit. The arithmetic means for both varieties, the significance level, and the p-value are presented.
Fibers 07 00064 g011
Figure 12. (a) Surface deformation images of the fiber samples obtained in the PeakForce tapping mode; (b) the surface deformation distributions and their Gaussian fit. The arithmetic means for both varieties, the significance level, and the p-value are presented.
Figure 12. (a) Surface deformation images of the fiber samples obtained in the PeakForce tapping mode; (b) the surface deformation distributions and their Gaussian fit. The arithmetic means for both varieties, the significance level, and the p-value are presented.
Fibers 07 00064 g012
Figure 13. Theoretical contact radius a as a function of normal force L and adhesion energy for a silicon nitride tip with a radius of curvature of 40 nm, obtained using Hertz, Derjaguin–Muller–Toporov (DMT), and Jonson–Kendall–Roberts (JKR) contact mechanics models. Each line represents the variation in the real contact area for a given adhesion energy, ranging from 0 to 30 N/m.
Figure 13. Theoretical contact radius a as a function of normal force L and adhesion energy for a silicon nitride tip with a radius of curvature of 40 nm, obtained using Hertz, Derjaguin–Muller–Toporov (DMT), and Jonson–Kendall–Roberts (JKR) contact mechanics models. Each line represents the variation in the real contact area for a given adhesion energy, ranging from 0 to 30 N/m.
Fibers 07 00064 g013
Table 1. Physical and dimensional properties of cotton fiber samples.
Table 1. Physical and dimensional properties of cotton fiber samples.
SamplesMacroscale Friction Coefficient MicronaireLength [cm]Strength [g/tex]Elongation [%]+b
A0.0804.72.832.17.48.4
B0.0664.12.833.18.69.6

Share and Cite

MDPI and ACS Style

Hosseinali, F.; Thomasson, J.A. Probing of Nanoscale Friction and Mechanical Characteristics of Cotton Fiber’s Surface. Fibers 2019, 7, 64. https://0-doi-org.brum.beds.ac.uk/10.3390/fib7070064

AMA Style

Hosseinali F, Thomasson JA. Probing of Nanoscale Friction and Mechanical Characteristics of Cotton Fiber’s Surface. Fibers. 2019; 7(7):64. https://0-doi-org.brum.beds.ac.uk/10.3390/fib7070064

Chicago/Turabian Style

Hosseinali, Farzad, and John Alexander Thomasson. 2019. "Probing of Nanoscale Friction and Mechanical Characteristics of Cotton Fiber’s Surface" Fibers 7, no. 7: 64. https://0-doi-org.brum.beds.ac.uk/10.3390/fib7070064

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

Article Metrics

Back to TopTop