Jump to content

User:Cdnbyu/sandbox

From Wikipedia, the free encyclopedia

Surface Force Apparatus[edit]

[1] [2] [3] [4] [5][6][7][8]

The surface force apparatus (SFA) is a scientific instrument which measures the interaction force of two surfaces as they are brought together and retracted. One surface is held by a cantilevered spring, and the deflection of the spring is used to calculate the force being exerted.[9] The technique was pioneered by D. Tabor, R.H.S. Winterton, J.N. Israelachvili in the early 1970s at Cambridge University. By the mid-1970s Israelachvili had adapted the original design to operate in liquids, notably aqueous solutions, while at the Australian National University.

Theory[edit]

The force measurements of the SFA are based primarily on Hooke's Law,

where F is the restoring force of a spring, k is the spring constant and x is the displacement of the spring.

Using a cantilevered spring, the lower surface is brought towards the top surface using a fine micrometer or piezo tube. The force between the two surfaces is measured by

where is the change in displacement applied by the micrometer and is the change displacement measured by interferometry.

The spring constants can range anywhere from to .

Operation[edit]

A surface force apparatus uses piezoelectric positioning elements (in addition to conventional motors for coarse adjustments), and senses the distance between the surfaces using optical interferometry.[citation needed] Using these sensitive elements, the device can resolve distances to within 0.1 nanometer, and forces at the 10−8 N level. This extremely sensitive technique can be used to measure electrostatic forces, elusive van der Waals forces, and even hydration or solvation forces. SFA is in some ways similar to using an atomic force microscope to measure interaction between a tip (or molecule adsorbed onto the tip) and a surface. The SFA, however, is more ideally suited to measuring surface-surface interactions, and can measure much longer-range forces more accurately. The SFA technique is quite demanding, however, and only a handful of labs worldwide have functional instruments.

In the SFA method, two smooth cylindrically curved surfaces whose cylindrical axes are positioned at 90° to each other are made to approach each other in a direction normal to the axes. The distance between the surfaces at the point of closest approach varies between a few micrometers to a few nanometers depending on the apparatus. When the two curved cylinders have the same radius of curvature, R, this so-called 'crossed cylinders' geometry is mathematically equivalent to the interaction between a flat surface and a sphere of radius R. Using the crossed cylinder geometry makes alignment much easier, enables testing of many different surface regions for better statistics, and also enables angle-dependent measurements to be taken. A typical setup involves R = 1 cm.

An example SFA setup to show the various layers, using a geometrically equivalent model

NEW Paragraph

Position measurements are typically made using multiple beam interferometry (MBI). The transparent surfaces of the perpendicular cylinders, usually mica, are backed with a highly reflective material usually silver before being mounted to the glass cylinders. When a white-light source is shined normal to the perpendicular cylinders the light will reflect back and forth until it is transmitted at where the surfaces are closest. These rays create an interference pattern, known as fringes of equal chromatic order (FECO), which can be observed by microscope. Distance between the two surfaces can be determined by analyzing these patterns. Mica is used because it is extremely flat, easy to work with, and optically transparent. Any other material or molecule of interest can be coated or adsorbed onto the mica layer.

The Jump Method[edit]

In the jump method, the top cylinder is mounted to a pair of cantilever springs, while the bottom cylinder is brought up towards the top cylinder. While the bottom cylinder approaches the top, there comes a point when they will "jump" into contact with each other. The measurements, in this case, are based the distance from which they jump and the spring constant. These measurements are usually between surfaces 1.25 nm and 20 nm apart. [10]

The Resonance Method[edit]

The jump method is difficult to execute mainly due to unaccounted vibrations entering the instrument. To overcome this, researchers developed the resonance method which measured surface forces at larger distances, 10 nm to 130 nm. In this case, the bottom cylinder is oscillated at a known frequency, while the frequency of the top cylinder is measured using a piezoelectric bimorph strain gauge. To minimize the dampening due to the surrounding substance, these measurements were originally done in a vacuum. [10]

Solvent mode[edit]

Early experiments measured the force between mica surfaces in air or vacuum.[10] The technique has been extended, however, to enable an arbitrary vapor or solvent to be introduced between the two surfaces.[11] In this way, interactions in various media can be carefully probed, and the dielectric constant of the gap between the surfaces can be tuned. Moreover, use of water as a solvent enables the measurement of interactions between biological molecules (such as lipids in biological membranes or proteins) in their native environment. In a solvent environment, SFA can even measure the oscillatory solvation and structural forces arising from the packing of individual layers of solvent molecules. It can also measure the electrostatic 'double layer' forces between charged surfaces in an aqueous medium with electrolyte.

Dynamic mode[edit]

The SFA has more recently been extended to perform dynamic measurements, thereby determining viscous and viscoelastic properties of fluids, frictional and tribological properties of surfaces, and the time-dependent interaction between biological structures.[12]

See also[edit]

References[edit]

  1. ^ Israelachvili, J; Min, Y; Akbulut, M; Alig, A; Carver, G; Greene, W; Kristiansen, K; Meyer, E; Pesika, N; Rosenberg, K; Zeng, H (1 March 2010). "Recent advances in the surface forces apparatus (SFA) technique". Reports on Progress in Physics. 73 (3): 036601. doi:10.1088/0034-4885/73/3/036601.
  2. ^ Wang, Yong Jian; Li, Feng; Rodriguez, Nicolas; Lafosse, Xavier; Gourier, Christine; Perez, Eric; Pincet, Frederic (29 March 2016). "Snapshot of sequential SNARE assembling states between membranes shows that N-terminal transient assembly initializes fusion". Proceedings of the National Academy of Sciences. 113 (13): 3533–3538. doi:10.1073/pnas.1518935113.
  3. ^ Perez, Eric; Li, Feng; Tareste, David; Pincet, Frédéric (10 October 2008). "The Surface Force Apparatus to Reveal the Energetics of Biomolecules Assembly. Application to DNA Bases Pairing and SNARE Fusion Proteins Folding". Cellular and Molecular Bioengineering. 1 (4): 240–246. doi:10.1007/s12195-008-0025-7.
  4. ^ Marra, Johan; Israelachvili, Jacob (August 1985). "Direct measurements of forces between phosphatidylcholine and phosphatidylethanolamine bilayers in aqueous electrolyte solutions". Biochemistry. 24 (17): 4608–4618. doi:10.1021/bi00338a020.
  5. ^ Helm, CA; Israelachvili, JN; McGuiggan, PM (17 November 1989). "Molecular mechanisms and forces involved in the adhesion and fusion of amphiphilic bilayers". Science (New York, N.Y.). 246 (4932): 919–22. PMID 2814514.
  6. ^ Craig, Vincent S. J. (1997-11-30). "An historical review of surface force measurement techniques". Colloids and Surfaces A: Physicochemical and Engineering Aspects. Prof. B. W. Ninham. 129: 75–93. doi:10.1016/S0927-7757(97)00029-0.
  7. ^ Marra, Johan; Israelachvili, Jacob (1985-08-01). "Direct measurements of forces between phosphatidylcholine and phosphatidylethanolamine bilayers in aqueous electrolyte solutions". Biochemistry. 24 (17): 4608–4618. doi:10.1021/bi00338a020. ISSN 0006-2960.
  8. ^ Helm, Christiane A.; Israelachvili, Jacob N.; McGuiggan, Patty M. (1992-02-01). "Role of hydrophobic forces in bilayer adhesion and fusion". Biochemistry. 31 (6): 1794–1805. doi:10.1021/bi00121a030. ISSN 0006-2960.
  9. ^ Israelachvili, J; Min, Y; Akbulut, M; Alig, A; Carver, G; Greene, W; Kristiansen, K; Meyer, E; Pesika, N; Rosenberg, K; Zeng, H (2010). "Recent advances in the surface forces apparatus (SFA) technique". Reports on Progress in Physics. 73 (3): 036601. Bibcode:2010RPPh...73c6601I. doi:10.1088/0034-4885/73/3/036601. ISSN 0034-4885.
  10. ^ a b c Israelachvili, J. N.; Tabor, D. (1972-11-21). "The Measurement of Van Der Waals Dispersion Forces in the Range 1.5 to 130 nm". Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences. 331 (1584): 19–38. doi:10.1098/rspa.1972.0162. ISSN 1364-5021.
  11. ^ Israelachvili, J. N.; Adams, G. E. (1976-08-26). "Direct measurement of long range forces between two mica surfaces in aqueous KNO3 solutions". Nature. 262 (5571): 774–776. doi:10.1038/262774a0.
  12. ^ A new surface forces apparatus for nanorheology, Review of Scientific Instruments 73 (6), 2296 (2002). DOI: 10.1063/1.1476719

Further reading[edit]

Evolution of Echolocation[edit]

[1] [2][3]

Social Effect of Diabetes[edit]

[4][5]

Also add you article title to the dashboard listing please. DJW56 (talk) 05:29, 9 October 2016 (UTC)

  1. ^ Parker, Joe; Tsagkogeorga, Georgia; Cotton, James A.; Liu, Yuan; Provero, Paolo; Stupka, Elia; Rossiter, Stephen J. (4 September 2013). "Genome-wide signatures of convergent evolution in echolocating mammals". Nature. 502 (7470): 228–231. doi:10.1038/nature12511.
  2. ^ Herron, Scott Freeman, Jon C. (2012). Evolutionary analysis (5th ed. ed.). San Francisco, Calif.: Benjamin Cummings. ISBN 0321616677. {{cite book}}: |edition= has extra text (help)CS1 maint: multiple names: authors list (link)
  3. ^ Liu, Zhen; Qi, Fei-Yan; Zhou, Xin; Ren, Hai-Qing; Shi, Peng (2014-09-01). "Parallel Sites Implicate Functional Convergence of the Hearing Gene Prestin among Echolocating Mammals". Molecular Biology and Evolution. 31 (9): 2415–2424. doi:10.1093/molbev/msu194. ISSN 0737-4038. PMID 24951728.
  4. ^ Hill, Jacqueline; Nielsen, Marcia; Fox, Michael H (2013-01-01). "Understanding the Social Factors That Contribute to Diabetes: A Means to Informing Health Care and Social Policies for the Chronically Ill". The Permanente Journal. 17 (2): 67–72. doi:10.7812/TPP/12-099. ISSN 1552-5767. PMC 3662286. PMID 23704847.
  5. ^ Gross, Lee S.; Li, Li; Ford, Earl S.; Liu, Simin (2004-05-01). "Increased consumption of refined carbohydrates and the epidemic of type 2 diabetes in the United States: an ecologic assessment". The American Journal of Clinical Nutrition. 79 (5): 774–779. ISSN 0002-9165. PMID 15113714.