Summary

Atomic Force Microscopy Cantilever-Based Nanoindentation: Mechanical Property Measurements at the Nanoscale in Air and Fluid

Published: December 02, 2022
doi:

Summary

Quantifying the contact area and force applied by an atomic force microscope (AFM) probe tip to a sample surface enables nanoscale mechanical property determination. Best practices to implement AFM cantilever-based nanoindentation in air or fluid on soft and hard samples to measure elastic modulus or other nanomechanical properties are discussed.

Abstract

An atomic force microscope (AFM) fundamentally measures the interaction between a nanoscale AFM probe tip and the sample surface. If the force applied by the probe tip and its contact area with the sample can be quantified, it is possible to determine the nanoscale mechanical properties (e.g., elastic or Young's modulus) of the surface being probed. A detailed procedure for performing quantitative AFM cantilever-based nanoindentation experiments is provided here, with representative examples of how the technique can be applied to determine the elastic moduli of a wide variety of sample types, ranging from kPa to GPa. These include live mesenchymal stem cells (MSCs) and nuclei in physiological buffer, resin-embedded dehydrated loblolly pine cross-sections, and Bakken shales of varying composition.

Additionally, AFM cantilever-based nanoindentation is used to probe the rupture strength (i.e., breakthrough force) of phospholipid bilayers. Important practical considerations such as method choice and development, probe selection and calibration, region of interest identification, sample heterogeneity, feature size and aspect ratio, tip wear, surface roughness, and data analysis and measurement statistics are discussed to aid proper implementation of the technique. Finally, co-localization of AFM-derived nanomechanical maps with electron microscopy techniques that provide additional information regarding elemental composition is demonstrated.

Introduction

Understanding the mechanical properties of materials is one of the most fundamental and essential tasks in engineering. For the analysis of bulk material properties, there are numerous methods available to characterize the mechanical properties of material systems, including tensile tests1, compression tests2, and three- or four-point bending (flexural) tests3. While these microscale tests can provide invaluable information regarding bulk material properties, they are generally conducted to failure, and are hence destructive. Additionally, they lack the spatial resolution necessary to accurately investigate the micro- and nanoscale properties of many material systems that are of interest today, such as thin films, biological materials, and nanocomposites. To begin addressing some of the problems with large-scale mechanical testing, mainly its destructive nature, microhardness tests were adopted from mineralogy. Hardness is a measure of the resistance of a material to plastic deformation under specific conditions. In general, microhardness tests use a stiff probe, usually made from hardened steel or diamond, to indent into a material. The resulting indentation depth and/or area can then be used to determine the hardness. Several methods have been developed, including Vickers4, Knoop5, and Brinell6 hardness; each provides a measure of microscale material hardness, but under different conditions and definitions, and as such only produces data that can be compared to tests performed under the same conditions.

Instrumented nanoindentation was developed to improve upon the relative values obtained via the various microhardness testing methods, improve the spatial resolution possible for the analysis of mechanical properties, and enable the analysis of thin films. Importantly, by utilizing the method first developed by Oliver and Pharr7, the elastic or Young's modulus, E, of a sample material can be determined via instrumented nanoindentation. Furthermore, by employing a Berkovich three-sided pyramidal nanoindenter probe (whose ideal tip area function matches that of the Vickers four-sided pyramidal probe)8, direct comparison between nanoscale and more traditional microscale hardness measurements can be made. With the growth in popularity of the AFM, AFM cantilever-based nanoindentation began receiving attention as well, particularly for measuring the mechanical properties of softer materials. As a result, as depicted schematically in Figure 1, the two most commonly employed techniques today to interrogate and quantify nanoscale mechanical properties are instrumented nanoindentation (Figure 1A) and AFM cantilever-based nanoindentation (Figure 1B)9, the latter of which is the focus of this work.

Figure 1
Figure 1: Comparison of instrumented and AFM cantilever-based nanoindentation systems. Schematic diagrams depicting typical systems for conducting (A) instrumented nanoindentation and (B) AFM cantilever-based nanoindentation. This figure was modified from Qian et al.51. Abbreviation: AFM = atomic force microscopy. Please click here to view a larger version of this figure.

Both instrumented and AFM cantilever-based nanoindentation employ a stiff probe to deform a sample surface of interest and monitor the resultant force and displacement as a function of time. Typically, either the desired load (i.e., force) or (Z-piezo) displacement profile is specified by the user via the software interface and directly controlled by the instrument, while the other parameter is measured. The mechanical property most often obtained from nanoindentation experiments is the elastic modulus (E), also referred to as the Young's modulus, which has units of pressure. The elastic modulus of a material is a fundamental property relating to the bond stiffness and is defined as the ratio of tensile or compressive stress (σ, the applied force per unit area) to axial strain (ε, the proportional deformation along the indentation axis) during elastic (i.e., reversible or temporary) deformation prior to the onset of plastic deformation (equation [1]):

Equation 1    (1)

It should be noted that, because many materials (especially biological tissues) are in fact viscoelastic, in reality, the (dynamic or complex) modulus consists of both elastic (storage, in phase) and viscous (loss, out of phase) components. In actual practice, what is measured in a nanoindentation experiment is the reduced modulus, E*, which is related to the true sample modulus of interest, E, as shown in equation (2):

Equation 2    (2)

Where Etip and νtip are the elastic modulus and Poisson's ratio, respectively, of the nanoindenter tip, and ν is the estimated Poisson's ratio of the sample. The Poisson's ratio is the negative ratio of the transverse to axial strain, and hence indicates the degree of transverse elongation of a sample upon being subjected to axial strain (e.g., during nanoindentation loading), as shown in equation (3):

Equation 3    (3)

The conversion from reduced to actual modulus is necessary because a) some of the axial strain imparted by the indenter tip may be converted to transverse strain (i.e., the sample may deform via expansion or contraction perpendicular to the direction of loading), and b) the indenter tip is not infinitely hard, and thus the act of indenting the sample results in some (small) amount of deformation of the tip. Note that in the case where Etip >> E (i.e., the indenter tip is much harder than the sample, which is often true when using a diamond probe), the relationship between the reduced and actual sample modulus simplifies greatly to EE*(1 - v2). While instrumented nanoindentation is superior in terms of accurate force characterization and dynamic range, AFM cantilever-based nanoindentation is faster, provides orders of magnitude greater force and displacement sensitivity, enables higher resolution imaging and improved indentation locating, and can simultaneously probe nanoscale magnetic and electrical properties9. In particular, AFM cantilever-based nanoindentation is superior for the quantification of mechanical properties at the nanoscale of soft materials (e.g., polymers, gels, lipid bilayers, and cells or other biological materials), extremely thin (sub-µm) films (where substrate effects can come into play depending upon indentation depth)10,11, and suspended two-dimensional (2D) materials12,13,14 such as graphene15,16, mica17, hexagonal boron nitride (h-BN)18, or transition metal dichalcogenides (TMDCs; e.g., MoS2)19. This is due to its exquisite force (sub-nN) and displacement (sub-nm) sensitivity, which is important for accurately determining the initial point of contact and remaining within the elastic deformation region.

In AFM cantilever-based nanoindentation, displacement of an AFM probe toward the sample surface is actuated by a calibrated piezoelectric element (Figure 1B), with the flexible cantilever eventually bending due to the resistive force experienced upon contact with the sample surface. This bending or deflection of the cantilever is typically monitored by reflecting a laser off the back of the cantilever and into a photodetector (position sensitive detector [PSD]). Coupled with the knowledge of the cantilever stiffness (in nN/nm) and deflection sensitivity (in nm/V), it is possible to convert this measured cantilever deflection (in V) into the force (in nN) applied to the sample. Following contact, the difference between the Z-piezo movement and the cantilever deflection yields the sample indentation depth. Combined with the knowledge of the tip area function, this enables calculation of the tip-sample contact area. The slope of the in-contact portions of the resulting force-distance or force-displacement (F-D) curves can then be fit using an appropriate contact mechanics model (see the Data Analysis section of the discussion) to determine the nanomechanical properties of the sample. While AFM cantilever-based nanoindentation possesses some distinct advantages over instrumented nanoindentation as described above, it also presents several practical implementation challenges, such as calibration, tip wear, and data analysis, which will be discussed here. Another potential downside of AFM cantilever-based nanoindentation is the assumption of linear elasticity, as the contact radius and indentation depths need to be much smaller than the indenter radius, which can be difficult to achieve when working with nanoscale AFM probes and/or samples exhibiting significant surface roughness.

Traditionally, nanoindentation has been limited to individual locations or small grid indentation experiments, wherein a desired location (i.e., region of interest [ROI]) is selected and a single controlled indent, multiple indents in a single location separated by some waiting time, and/or a coarse grid of indents are performed at a rate on the order of Hz. However, recent advances in AFM allow for the simultaneous acquisition of mechanical properties and topography through the utilization of high-speed force curve-based imaging modes (referred to by various tradenames depending on the system manufacturer), wherein force curves are conducted at a kHz rate under load control, with the maximum tip-sample force utilized as the imaging setpoint. Point-and-shoot methods have also been developed, allowing for the acquisition of an AFM topography image followed by subsequent selective nanoindentation at points of interest within the image, affording nanoscale spatial control over nanoindentation location. While not the primary focus of this work, specific selected application examples of both force curve-based imaging and point-and-shoot cantilever-based nanoindentation are presented in the representative results, and can be used in conjunction with the protocol outlined below if available on the particular AFM platform employed. Specifically, this work outlines a generalized protocol for the practical implementation of AFM cantilever-based nanoindentation on any capable AFM system and provides four use case examples (two in air, two in fluid) of the technique, including representative results and an in-depth discussion of the nuances, challenges, and important considerations to successfully employ the technique.

Protocol

NOTE: Due to the wide variety of commercially available AFMs and diversity of sample types and applications that exist for cantilever-based nanoindentation, the protocol that follows is intentionally designed to be relatively general in nature, focusing on the shared steps necessary for all cantilever-based nanoindentation experiments regardless of instrument or manufacturer. Because of this, the authors assume the reader possesses at least basic familiarity with operating the specific instrument chosen for performing ca…

Representative Results

Force-displacement curves Figure 7 shows representative, near-ideal F-D curves obtained from nanoindentation experiments performed in air on resin-embedded loblolly pine samples (Figure 7A) and in fluid (phosphate-buffered saline [PBS]) on mesenchymal stem cell (MSC) nuclei (Figure 7B). The use of any contact mechanics model relies on the accurate and reliable determination of the initial tip-sample contact po…

Discussion

Sample preparation
For nanoindentation in air, common preparation methods include cryosectioning (e.g., tissue samples), grinding and/or polishing followed by ultramicrotoming (e.g., resin-embedded biological samples), ion milling or focused ion beam preparation (e.g., semiconductor, porous, or mixed hardness samples not amenable to polishing), mechanical or electrochemical polishing (e.g., metal alloys), or thin film deposition (e.g., atomic layer or chemical vapor deposition, molecular beam epita…

Divulgations

The authors have nothing to disclose.

Acknowledgements

All AFM experiments were performed in the Boise State University Surface Science Laboratory (SSL). SEM characterization was performed in the Boise State Center for Materials Characterization (BSCMC). Research reported in this publication regarding biofuel feedstocks was supported in part by the US Department of Energy, Office of Energy Efficiency and Renewable Energy, Bioenergy Technologies Office as part of the Feedstock Conversion Interface Consortium (FCIC), and under DOE Idaho Operations Office Contract DE-AC07-051ID14517. Cell mechanics studies were supported by the National Institutes of Health (USA) under grants AG059923, AR075803, and P20GM109095, and by National Science Foundation (USA) grants 1929188 and 2025505. The model lipid bilayer systems work was supported by the National Institutes of Health (USA) under grant R01 EY030067. The authors thank Dr. Elton Graugnard for producing the composite image shown in Figure 11.

Materials

Atomic force microscope Bruker Dimension Icon Uses Nanoscope control software, including PeakForce Quantitative Nanomechanical Mapping (PF-QNM), FastForce Volume (FFV), and Point-and-Shoot Ramping experimental workspaces
AtomicJ American Institute of Physics https://doi.org/10.1063/1.4881683 Flexible, powerful, free open source Java-based force curve analysis software package. Supports numerous contact mechanic models, such as Hertz, Sneddon DMT, JKR, Maugis, and cone or pyramid (including blunt and truncated). Also includes a variety of initial contact point estimation methods to choose from. Supports batch processing of data and subsequent statistical analysis (e.g., averages, standard deviations, histograms, goodness of fit, etc.). Literature citation is: P. Hermanowicz, M. Sarna, K. Burda, and H. GabryEquation 1, “AtomicJ: An open source software for analysis of force curves” Rev. Sci. Instrum. 85: 063703 (2014), https://doi.org/10.1063/1.4881683
Buffer solution (PBS) Fisher Chemical (NaCl), Sigma Aldrich (KCl), Fisher BioReagents (Na2HPO4 and KH2PO4) S271 (>99% purity NaCl), P9541 (>99% purity KCl), BP332(>99% purity Na2HPO4), BP362 (>99% purity KH2PO4) Phosphate buffered saline (PBS) was prepared in the laboratory as an aqueous solution consisting of 137 mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, and 1.8 mM KH2PO4 dissolved in ultrapure water. Reagents were measured out using an analytical balance, and glassware was cleaned with soap and water followed by autoclaving immediately prior to use.
Chloroform
Diamond tip AFM probe Bruker PDNISP Pre-mounted factory-calibrated cube corner diamond (E = 1140 GPa) tip AFM probe (nominal R = 40 nm) with a stainless steel cantilever (nominal k = 225 N/m, f0 = 50 kHz). Spring constant is measured at the factory (k = 256 N/m for the probe, Serial #13435414, used here) and calibration data (including AFM images of indents showing probe geometry) is provided with the probe.
Diamond ultramicrotome blade Diatome Ultra 35° 2.1 mm width. Also used a standard glass blade for intial rough cut of sample surface before transitioning to diamond blade for final surface preparation
Epoxy Gorilla Glue 26853-31-6 Epoxy resin and hardner were mixed in a 1:1 ratio, a small drop was placed on a stainless steel sample puck (Ted Pella), and V1 grade muscovite mica (Ted Pella) was attached to create an atomically flat surface for preparation of phospholipid membranes.
Ethanol
LR white resin, medium grade (catalyzed) Electron Microscopy Sciences 14381 500 mL bottle, Lot #150629
Mesenchymal stem cells (MSCs) N/A N/A MSCs for nanomechanical studies were primary cells harvested from 8-10 week old male C57BL/6 mice as described in Goelzer, M. et al. "Lamin A/C Is Dispensable to Mechanical Repression of Adipogenesis" Int J Mol Sci 22: 6580 (2021) doi:10.3390/ijms22126580 and Peister, A. et al. "Adult stem cells from bone marrow (MSCs) isolated from different strains of inbred mice vary in surface epitopes, rates of proliferation, and differentiation potential" Blood 103: 1662-1668 (2004), doi:10.1182/blood-2003-09-3070.
Modulus standards Bruker PFQNM-SMPKIT-12M Used HOPG (E = 18 GPa) and PS (E = 2.7 GPa). Also contains 2x PDMS (Tack 0, E = 2.5 MPa; Tack 4, E = 3.5 MPa), PS-LDPE (E = 2.0/0.2 GPa), fused silica (E = 72.9 GPa), sapphire (E – 345 GPa), and tip characterization (titanium roughness) sample. All samples come pre-mounted on a 12 mm diameter steel disc (sample puck).
Muscovite mica Ted Pella 50-12 12 mm diameter, V1 grade muscovite mica
Nanscope Analysis Bruker Version 2.0 Free AFM image processing and analysis software package, but designed for, and proprietary/limited to Bruker AFMs; similar functionality is available from free, platform-independent AFM image processing and analysis software packages such as Gwyddion, WSxM, and others. Has built-in capabilities for force curve analysis, but AtomicJ is more flexible/full featured (e.g., more built-in contact mechanics models to choose from, statistical analysis of force curve fitting results, etc.) for force curve analysis and handles batch processing of force curves.
Phospholipids: POPC, Cholesterol (ovine) Avanti Polar Lipids POPC: CAS # 26853-31-6, Cholesterol: CAS # 57-88-5 POPC lipid dissolved in chloroform (25 mg/mL) was obtained from vendor and used without further purification. Cholesterol powder from the same vendor was dissolved in chloroform (20 mg/mL). 
Probe holder (fluid, lipid bilayers) Bruker MTFML-V2 Specific to the particular AFM used; MTFML-V2 is a glass probe holder for scanning in fluid on a MultiMode AFM.
Probe holder (fluid, MSCs) Bruker FastScan Bio Z-scanner Used with Dimension FastScan head (XY flexure scanners). Serial number MXYPOM5-1B154.
Probe holder (standard, ambient) Bruker DAFMCH Specific to the particular AFM used; DAFMCH is the standard contact and tapping mode probe holder for the Dimension Icon AFM, suitable for nanoindentation (PF-QNM, FFV, and point-and-shoot ramping)
Sample Puck Ted Pella 16218 Product number is for 15 mm diameter stainless steel sample puck. Also available in 6 mm, 10 mm, 12 mm, and 20 mm diameters at https://www.tedpella.com/AFM_html/AFM.aspx#anchor842459
Sapphire substrate Bruker PFQNM-SMPKIT-12M Extremely hard surface (E = 345 GPa) for measuring deflection sensitivity of probes (want all of the deflection to come from the probe, not the substrate). Part of the PF-QNM/modulus standards kit.
Scanning electron microscope Hitachi S-3400N-II Located at Boise State. Used to perform co-localized SEM/EDS on all samples except additively manufactured (AM) Ti-6Al-4V.
Silicon AFM probes (standard) NuNano Scout 350 Standard tapping mode silicon probe with reflective aluminum backside coating; k = 42 N/m (nominal), f0 = 350 kHz. Nominal R = 5 nm. Also available uncoated or with reflective gold backside coating. Probes with similar specifications are available from other manufacturers (e.g., Bruker TESPA-V2).
Silicon AFM probes (stiff) Bruker RTESPA-525, RTESPA-525-30  Rotated tip etched silicon probes with reflective aluminum backside coating; k = 200 N/m (nominal), f0 = 525 kHz. Nominal R = 8 nm for RTESPA-525, R = 30 nm for RTESPA-525-30. Spring constant of each RTESPA-525-30 is measured individually at the factory via laser Doppler vibrometry and supplied with the probe.
Silicon carbide grit paper (abrasive discs) Allied 50-10005 120 grit
Silicon nitride AFM probes (soft, large radius hemispherical tip) Bruker MLCT-SPH-5UM, MLCT-SPH-5UM-DC Also MLCT-SPH-1UM-DC. New product line of factory-calibrated (probe radius and spring constants of all cantilevers) large radius (R = 1 or 5 mm) hemispherical tip (at the end of a 23 mm long cylindrical shaft) probes. DC = drift compensation coating. 6 cantilevers/probe (A-F). Nominal spring constants: A, k = 0.07 N/m; B, k = 0.02 N/m; C, k = 0.01 N/m; D, k = 0.03 N/m; E, k = 0.1 N/m; F, k = 0.6 N/m.
Silicon nitride AFM probes (soft, medium sharp tip) Bruker DNP 4 cantilevers/probe (A-d). Nominal spring constants: A, k = 0.35 N/m; B, k = 0.12 N/m; C, k = 0.24 N/m; D, k = 0.06 N/m. Nominal radii of curvature, R = 10 nm.
Silicon nitride AFM probes (soft, sharp tip) Bruker ScanAsyst-Air Nominal values: resonance frequency, f0 = 70 kHz; spring constant, k = 0.4 N/m; radius of curvature, R = 2 nm. Designed for force curve based AFM imaging.
Superglue Henkel Loctite 495 Cyanoacrylate based instant adhesive. Lots of roughly equivalent products are readily available.
Syringe pump New Era Pump Systems NE1000US One channel syringe pump system with infusion and withdrawal capacity
Tip characterization standard Bruker PFQNM-SMPKIT-12M Titanium (Ti) roughness standard. Part of the PF-QNM/modulus standards kit.
Ultrahigh purity nitrogen (UHP N2), 99.999% Norco SPG TUHPNI – T T size compressed gas cylinder of ultrahigh purity (99.999%) nitrogen for drying samples
Ultramicrotome Leica EM UC6 Equipped with a glass blade (standard, for intial sample preparation) and a diamond blade (for final preparation)
Ultrapure water Thermo Fisher Barnstead Nanopure Model 7146 Model has been discontinued, but equivalent products are available. Produces ≥18.2 MΩ*cm ultrapure water with 1-5 ppb TOC (total organic content), per inline UV monitoring. Includes 0.2 µm particulate filter, ion exchange columns, and UV oxidation chamber.
Variable Speed Grinder Buehler EcoMet 3000 Used with silicon carbide grit papers during hand polishing.
Vibration isolation table (active) Herzan TS-140 Used with Bruker MultiMode AFM. Sits on a TMC 65-531 vibration isolation table. Bruker Dimension Icon AFM utilizes strictly passive vibration isolation (comes from manufacturer with custom acoustic hood, air table, and granite slab).
Vibration isolation table (passive) TMC 65-531 35" x 30" vibration isolation table with optional air damping (disabled). Used with Bruker MultiMode AFM. Herzan TS-140 "Table Stable" active vibration control table is located on top.

References

  1. Hart, E. W. Theory of the tensile test. Acta Metallurgica. 15 (2), 351-355 (1967).
  2. Fell, J. T., Newton, J. M. Determination of tablet strength by the diametral-compression test. Journal of Pharmaceutical Sciences. 59 (5), 688-691 (1970).
  3. Babiak, M., Gaff, M., Sikora, A., Hysek, &. #. 3. 5. 2. ;. Modulus of elasticity in three- and four-point bending of wood. Composite Structures. 204, 454-465 (2018).
  4. Song, S., Yovanovich, M. M. Relative contact pressure-Dependence on surface roughness and Vickers microhardness. Journal of Thermophysics and Heat Transfer. 2 (1), 43-47 (1988).
  5. Hays, C., Kendall, E. G. An analysis of Knoop microhardness. Metallography. 6 (4), 275-282 (1973).
  6. Hill, R., Storåkers, B., Zdunek, A. B. A theoretical study of the Brinell hardness test. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences. 423 (1865), 301-330 (1989).
  7. Oliver, W. C., Pharr, G. M. An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. Journal of Materials Research. 7 (6), 1564-1583 (1992).
  8. Sakharova, N. A., Fernandes, J. V., Antunes, J. M., Oliveira, M. C. Comparison between Berkovich, Vickers and conical indentation tests: A three-dimensional numerical simulation study. International Journal of Solids and Structures. 46 (5), 1095-1104 (2009).
  9. Cohen, S. R., Kalfon-Cohen, E. Dynamic nanoindentation by instrumented nanoindentation and force microscopy: a comparative review. Beilstein Journal of Nanotechnology. 4 (1), 815-833 (2013).
  10. Saha, R., Nix, W. D. Effects of the substrate on the determination of thin film mechanical properties by nanoindentation. Acta Materialia. 50 (1), 23-38 (2002).
  11. Tsui, T. Y., Pharr, G. M. Substrate effects on nanoindentation mechanical property measurement of soft films on hard substrates. Journal of Materials Research. 14 (1), 292-301 (1999).
  12. Cao, G., Gao, H. Mechanical properties characterization of two-dimensional materials via nanoindentation experiments. Progress in Materials Science. 103, 558-595 (2019).
  13. Castellanos-Gomez, A., Singh, V., vander Zant, H. S. J., Steele, G. A. Mechanics of freely-suspended ultrathin layered materials. Annalen der Physik. 527 (1-2), 27-44 (2015).
  14. Cao, C., Sun, Y., Filleter, T. Characterizing mechanical behavior of atomically thin films: A review. Journal of Materials Research. 29 (3), 338-347 (2014).
  15. Lee, C., Wei, X., Kysar, J. W., Hone, J. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science. 321 (5887), 385-388 (2008).
  16. Elibol, K., et al. Visualising the strain distribution in suspended two-dimensional materials under local deformation. Scientific Reports. 6 (1), 28485 (2016).
  17. Castellanos-Gomez, A., et al. Mechanical properties of freely suspended atomically thin dielectric layers of mica. Nano Research. 5 (8), 550-557 (2012).
  18. Song, L., et al. Large scale growth and characterization of atomic hexagonal boron nitride layers. Nano Letters. 10 (8), 3209-3215 (2010).
  19. Castellanos-Gomez, A., et al. Elastic properties of freely suspended MoS2 nanosheets. Advanced Materials. 24 (6), 772-775 (2012).
  20. D’Costa, N. P., Hoh, J. H. Calibration of optical lever sensitivity for atomic force microscopy. Review of Scientific Instruments. 66 (10), 5096-5097 (1995).
  21. Wu, Y., et al. Evaluation of elastic modulus and hardness of crop stalks cell walls by nano-indentation. Bioresource Technology. 101 (8), 2867-2871 (2010).
  22. Barns, S., et al. Investigation of red blood cell mechanical properties using AFM indentation and coarse-grained particle method. BioMedical Engineering OnLine. 16 (1), 140 (2017).
  23. Hermanowicz, P., Sarna, M., Burda, K., Gabryś, H. AtomicJ: An open source software for analysis of force curves. Review of Scientific Instruments. 85 (6), 063703 (2014).
  24. Broitman, E. Indentation hardness measurements at macro-, micro-, and nanoscale: a critical overview. Tribology Letters. 65 (1), 23 (2016).
  25. Tiwari, A. . Nanomechanical Analysis of High Performance Materials. , (2015).
  26. Aggarwal, R. L., Ramdas, A. K. . Physical Properties of Diamond and Sapphire. , (2019).
  27. Boyd, E. J., Uttamchandani, D. Measurement of the anisotropy of Young’s modulus in single-crystal silicon. Journal of Microelectromechanical Systems. 21 (1), 243-249 (2012).
  28. Harding, J. W., Sneddon, I. N. The elastic stresses produced by the indentation of the plane surface of a semi-infinite elastic solid by a rigid punch. Mathematical Proceedings of the Cambridge Philosophical Society. 41 (1), 16-26 (2008).
  29. Lin, D. C., Dimitriadis, E. K., Horkay, F. Robust strategies for automated AFM force curve analysis-I. Non-adhesive indentation of soft, inhomogeneous materials. Journal of Biomechanical Engineering. 129 (3), 430-440 (2006).
  30. Lin, D. C., Dimitriadis, E. K., Horkay, F. Robust strategies for automated AFM force curve analysis-II: Adhesion-influenced indentation of soft, elastic materials. Journal of Biomechanical Engineering. 129 (6), 904-912 (2007).
  31. Haile, S., Palmer, M., Otey, A. Potential of loblolly pine: switchgrass alley cropping for provision of biofuel feedstock. Agroforestry Systems. 90 (5), 763-771 (2016).
  32. Lu, X., et al. Biomass logistics analysis for large scale biofuel production: Case study of loblolly pine and switchgrass. Bioresource Technology. 183, 1-9 (2015).
  33. Susaeta, A., Lal, P., Alavalapati, J., Mercer, E., Carter, D. Economics of intercropping loblolly pine and switchgrass for bioenergy markets in the southeastern United States. Agroforestry Systems. 86 (2), 287-298 (2012).
  34. Garcia, R. Nanomechanical mapping of soft materials with the atomic force microscope: methods, theory and applications. Chemical Society Reviews. 49 (16), 5850-5884 (2020).
  35. Derjaguin, B. V., Muller, V. M., Toporov, Y. P. Effect of contact deformations on the adhesion of particles. Journal of Colloid and Interface Science. 53 (2), 314-326 (1975).
  36. Ciesielski, P. N., et al. Engineering plant cell walls: tuning lignin monomer composition for deconstructable biofuel feedstocks or resilient biomaterials. Green Chemistry. 16 (5), 2627-2635 (2014).
  37. Liu, K., Ostadhassan, M., Zhou, J., Gentzis, T., Rezaee, R. Nanoscale pore structure characterization of the Bakken shale in the USA. Fuel. 209, 567-578 (2017).
  38. Maryon, O. O., et al. Co-localizing Kelvin probe force microscopy with other microscopies and spectroscopies: selected applications in corrosion characterization of alloys. JoVE. (184), e64102 (2022).
  39. Eliyahu, M., Emmanuel, S., Day-Stirrat, R. J., Macaulay, C. I. Mechanical properties of organic matter in shales mapped at the nanometer scale. Marine and Petroleum Geology. 59, 294-304 (2015).
  40. Li, C., et al. Nanomechanical characterization of organic matter in the Bakken formation by microscopy-based method. Marine and Petroleum Geology. 96, 128-138 (2018).
  41. Bouzid, T., et al. The LINC complex, mechanotransduction, and mesenchymal stem cell function and fate. Journal of Biological Engineering. 13 (1), 68 (2019).
  42. Dupont, S., et al. Role of YAP/TAZ in mechanotransduction. Nature. 474 (7350), 179-183 (2011).
  43. Wang, S., et al. CCM3 is a gatekeeper in focal adhesions regulating mechanotransduction and YAP/TAZ signalling. Nature Cell Biology. 23 (7), 758-770 (2021).
  44. Sen, B., et al. Mechanical strain inhibits adipogenesis in mesenchymal stem cells by stimulating a durable β-catenin signal. Endocrinology. 149 (12), 6065-6075 (2008).
  45. Sen, B., et al. mTORC2 regulates mechanically induced cytoskeletal reorganization and lineage selection in marrow-derived mesenchymal stem cells. Journal of Bone and Mineral Research. 29 (1), 78-89 (2014).
  46. Sen, B., et al. Mechanically induced nuclear shuttling of β-catenin requires co-transfer of actin. Stem Cells. 40 (4), 423-434 (2022).
  47. Newberg, J., et al. Isolated nuclei stiffen in response to low intensity vibration. Journal of Biomechanics. 111, 110012 (2020).
  48. Ding, Y., Xu, G. -. K., Wang, G. -. F. On the determination of elastic moduli of cells by AFM based indentation. Scientific Reports. 7 (1), 45575 (2017).
  49. Khadka, N. K., Timsina, R., Rowe, E., O’Dell, M., Mainali, L. Mechanical properties of the high cholesterol-containing membrane: An AFM study. Biochimica et Biophysica Acta. Biomembranes. 1863 (8), 183625 (2021).
  50. Castellana, E. T., Cremer, P. S. Solid supported lipid bilayers: From biophysical studies to sensor design. Surface Science Reports. 61 (10), 429-444 (2006).
  51. Qian, L., Zhao, H. Nanoindentation of soft biological materials. Micromachines. 9 (12), 654 (2018).
  52. Pittenger, B., Yablon, D. Improving the accuracy of nanomechanical measurements with force-curve-based AFM techniques. Bruker Application Notes. 149, (2017).
  53. Vorselen, D., Kooreman, E. S., Wuite, G. J. L., Roos, W. H. Controlled tip wear on high roughness surfaces yields gradual broadening and rounding of cantilever tips. Scientific Reports. 6 (1), 36972 (2016).
  54. Bhaskaran, H., et al. Ultralow nanoscale wear through atom-by-atom attrition in silicon-containing diamond-like carbon. Nature Nanotechnology. 5 (3), 181-185 (2010).
  55. Giannazzo, F., Schilirò, E., Greco, G., Roccaforte, F. Conductive atomic force microscopy of semiconducting transition metal dichalcogenides and heterostructures. Nanomaterials. 10 (4), 803 (2020).
  56. Melitz, W., Shen, J., Kummel, A. C., Lee, S. Kelvin probe force microscopy and its application. Surface Science Reports. 66 (1), 1-27 (2011).
  57. Kazakova, O., et al. Frontiers of magnetic force microscopy. Journal of Applied Physics. 125 (6), 060901 (2019).
  58. Kim, H. -. J., Yoo, S. -. S., Kim, D. -. E. Nano-scale wear: A review. International Journal of Precision Engineering and Manufacturing. 13 (9), 1709-1718 (2012).
  59. Heath, G. R., et al. Localization atomic force microscopy. Nature. 594 (7863), 385-390 (2021).
  60. Strahlendorff, T., Dai, G., Bergmann, D., Tutsch, R. Tip wear and tip breakage in high-speed atomic force microscopes. Ultramicroscopy. 201, 28-37 (2019).
  61. Lantz, M. A., et al. Wear-resistant nanoscale silicon carbide tips for scanning probe applications. Advanced Functional Materials. 22 (8), 1639-1645 (2012).
  62. Khurshudov, A. G., Kato, K., Koide, H. Wear of the AFM diamond tip sliding against silicon. Wear. 203, 22-27 (1997).
  63. Villarrubia, J. S. Algorithms for scanned probe microscope image simulation, surface reconstruction, and tip estimation. Journal of Research of the National Institute of Standards and Technology. 102 (4), 425 (1997).
  64. Kain, L., et al. Calibration of colloidal probes with atomic force microscopy for micromechanical assessment. Journal of the Mechanical Behavior of Biomedical Materials. 85, 225-236 (2018).
  65. Slattery, A. D., Blanch, A. J., Quinton, J. S., Gibson, C. T. Accurate measurement of Atomic Force Microscope cantilever deflection excluding tip-surface contact with application to force calibration. Ultramicroscopy. 131, 46-55 (2013).
  66. Dobrovinskaya, E. R., Lytvynov, L. A., Pishchik, V. . Sapphire: Material, Manufacturing, Applications. , (2009).
  67. te Riet, J., et al. Interlaboratory round robin on cantilever calibration for AFM force spectroscopy. Ultramicroscopy. 111 (12), 1659-1669 (2011).
  68. Pratt, J. R., Shaw, G. A., Kumanchik, L., Burnham, N. A. Quantitative assessment of sample stiffness and sliding friction from force curves in atomic force microscopy. Journal of Applied Physics. 107 (4), 044305 (2010).
  69. Slattery, A. D., Blanch, A. J., Quinton, J. S., Gibson, C. T. Calibration of atomic force microscope cantilevers using standard and inverted static methods assisted by FIB-milled spatial markers. Nanotechnology. 24 (1), 015710 (2012).
  70. Higgins, M. J., et al. Noninvasive determination of optical lever sensitivity in atomic force microscopy. Review of Scientific Instruments. 77 (1), 013701 (2006).
  71. Lévy, R., Maaloum, M. Measuring the spring constant of atomic force microscope cantilevers: thermal fluctuations and other methods. Nanotechnology. 13 (1), 33-37 (2001).
  72. Sikora, A. Quantitative normal force measurements by means of atomic force microscopy towards the accurate and easy spring constant determination. Nanoscience and Nanometrology. 2 (1), 8-29 (2016).
  73. Ohler, B. Cantilever spring constant calibration using laser Doppler vibrometry. Review of Scientific Instruments. 78 (6), 063701 (2007).
  74. Gates, R. S., Pratt, J. R. Accurate and precise calibration of AFM cantilever spring constants using laser Doppler vibrometry. Nanotechnology. 23 (37), 375702 (2012).
  75. Cleveland, J. P., Manne, S., Bocek, D., Hansma, P. K. A nondestructive method for determining the spring constant of cantilevers for scanning force microscopy. Review of Scientific Instruments. 64 (2), 403-405 (1993).
  76. Sader, J. E., Chon, J. W. M., Mulvaney, P. Calibration of rectangular atomic force microscope cantilevers. Review of Scientific Instruments. 70 (10), 3967-3969 (1999).
  77. Sader, J. E., et al. Spring constant calibration of atomic force microscope cantilevers of arbitrary shape. Review of Scientific Instruments. 83 (10), 103705 (2012).
  78. Sader, J. E. Frequency response of cantilever beams immersed in viscous fluids with applications to the atomic force microscope. Journal of Applied Physics. 84 (1), 64-76 (1998).
  79. Sader, J. E., Pacifico, J., Green, C. P., Mulvaney, P. General scaling law for stiffness measurement of small bodies with applications to the atomic force microscope. Journal of Applied Physics. 97 (12), 124903 (2005).
  80. Mendels, D. -. A., et al. Dynamic properties of AFM cantilevers and the calibration of their spring constants. Journal of Micromechanics and Microengineering. 16 (8), 1720-1733 (2006).
  81. Gao, S., Brand, U. In-situ nondestructive characterization of the normal spring constant of AFM cantilevers. Measurement Science and Technology. 25 (4), 044014 (2014).
  82. Gibson, C. T., Watson, G. S., Myhra, S. Determination of the spring constants of probes for force microscopy/spectroscopy. Nanotechnology. 7 (3), 259-262 (1996).
  83. Gates, R. S., Pratt, J. R. Prototype cantilevers for SI-traceable nanonewton force calibration. Measurement Science and Technology. 17 (10), 2852-2860 (2006).
  84. Neumeister, J. M., Ducker, W. A. Lateral, normal, and longitudinal spring constants of atomic force microscopy cantilevers. Review of Scientific Instruments. 65 (8), 2527-2531 (1994).
  85. Kim, M. S., Choi, I. M., Park, Y. K., Kang, D. I. Atomic force microscope probe calibration by use of a commercial precision balance. Measurement. 40 (7), 741-745 (2007).
  86. Kim, M. -. S., Choi, J. -. H., Park, Y. -. K., Kim, J. -. H. Atomic force microscope cantilever calibration device for quantified force metrology at micro- or nano-scale regime: the nano force calibrator (NFC). Metrologia. 43 (5), 389-395 (2006).
  87. Tian, Y., et al. A novel method and system for calibrating the spring constant of atomic force microscope cantilever based on electromagnetic actuation. Review of Scientific Instruments. 89 (12), 125119 (2018).
  88. Clifford, C. A., Seah, M. P. The determination of atomic force microscope cantilever spring constants via dimensional methods for nanomechanical analysis. Nanotechnology. 16 (9), 1666-1680 (2005).
  89. Chen, B. -. Y., Yeh, M. -. K., Tai, N. -. H. Accuracy of the spring constant of atomic force microscopy cantilevers by finite element method. Analytical Chemistry. 79 (4), 1333-1338 (2007).
  90. Mick, U., Eichhorn, V., Wortmann, T., Diederichs, C., Fatikow, S. Combined nanorobotic AFM/SEM system as novel toolbox for automated hybrid analysis and manipulation of nanoscale objects. 2010 IEEE International Conference on Robotics and Automation. , 4088-4093 (2010).
  91. Kim, M. -. S., Choi, J. -. H., Kim, J. -. H., Park, Y. -. K. Accurate determination of spring constant of atomic force microscope cantilevers and comparison with other methods. Measurement. 43 (4), 520 (2010).
  92. Zhang, G., Wei, Z., Ferrell, R. E. Elastic modulus and hardness of muscovite and rectorite determined by nanoindentation. Applied Clay Science. 43 (2), 271-281 (2009).
  93. Bobko, C. P., Ortega, J. A., Ulm, F. -. J. Comment on "Elastic modulus and hardness of muscovite and rectorite determined by nanoindentation by G. Zhang, Z. Wei and R.E. Ferrell. Applied Clay Science. 46 (4), 425-428 (2009).
  94. Zhang, G., Wei, Z., Ferrell, R. E. Reply to the Comment on "Elastic modulus and hardness of muscovite and rectorite determined by nanoindentation" by G. Zhang, Z. Wei and R. E. Ferrell. Applied Clay Science. 46 (4), 429-432 (2009).
  95. Jin, D. W., et al. Thermal stability and Young’s modulus of mechanically exfoliated flexible mica. Current Applied Physics. 18 (12), 1486-1491 (2018).
  96. Xiao, J., et al. Anisotropic friction behaviour of highly oriented pyrolytic graphite. Carbon. 65, 53-62 (2013).
  97. Hertz, H. Ueber die Berührung fester elastischer Körper. Journal für die reine und angewandte Mathematik. 1882 (92), 156-171 (1882).
  98. Johnson, K. L., Kendall, K., Roberts, A. D., Tabor, D. Surface energy and the contact of elastic solids. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences. 324 (1558), 301-313 (1971).
  99. Muller, V. M., Derjaguin, B. V., Toporov, Y. P. On two methods of calculation of the force of sticking of an elastic sphere to a rigid plane. Colloids and Surfaces. 7 (3), 251-259 (1983).
  100. Maugis, D. Adhesion of spheres: The JKR-DMT transition using a dugdale model. Journal of Colloid and Interface Science. 150 (1), 243-269 (1992).
  101. Muller, V. M., Yushchenko, V. S., Derjaguin, B. V. On the influence of molecular forces on the deformation of an elastic sphere and its sticking to a rigid plane. Journal of Colloid and Interface Science. 77 (1), 91-101 (1980).
  102. Muller, V. M., Yushchenko, V. S., Derjaguin, B. V. General theoretical consideration of the influence of surface forces on contact deformations and the reciprocal adhesion of elastic spherical particles. Journal of Colloid and Interface Science. 92 (1), 92-101 (1983).
  103. Johnson, K. L., Greenwood, J. A. An adhesion map for the contact of elastic spheres. Journal of Colloid and Interface Science. 192 (2), 326-333 (1997).
  104. Shi, X., Zhao, Y. -. P. Comparison of various adhesion contact theories and the influence of dimensionless load parameter. Journal of Adhesion Science and Technology. 18 (1), 55-68 (2004).
check_url/fr/64497?article_type=t

Play Video

Citer Cet Article
Enrriques, A. E., Howard, S., Timsina, R., Khadka, N. K., Hoover, A. N., Ray, A. E., Ding, L., Onwumelu, C., Nordeng, S., Mainali, L., Uzer, G., Davis, P. H. Atomic Force Microscopy Cantilever-Based Nanoindentation: Mechanical Property Measurements at the Nanoscale in Air and Fluid. J. Vis. Exp. (190), e64497, doi:10.3791/64497 (2022).

View Video