Microrheology

Last updated

Microrheology [1] is a technique used to measure the rheological properties of a medium, such as microviscosity, via the measurement of the trajectory of a flow tracer (a micrometre-sized particle). It is a new way of doing rheology, traditionally done using a rheometer. There are two types of microrheology: passive microrheology and active microrheology. Passive microrheology uses inherent thermal energy to move the tracers, whereas active microrheology uses externally applied forces, such as from a magnetic field or an optical tweezer, to do so. Microrheology can be further differentiated into 1- and 2-particle methods. [2] [3]

Contents

Passive microrheology

Passive microrheology uses the thermal energy (kT) to move the tracers, although recent evidence suggests that active random forces inside cells may instead move the tracers in a diffusive-like manner. [4] The trajectories of the tracers are measured optically either by microscopy, or alternatively by light scattering techniques. Diffusing-wave spectroscopy (DWS) is a common choice that extends light scattering measurement techniques to account for multiple scattering events. [5] From the mean squared displacement with respect to time (noted MSD or <Δr2> ), one can calculate the visco-elastic moduli G(ω) and G(ω) using the generalized Stokes–Einstein relation (GSER). Here is a view of the trajectory of a particle of micrometer size.

In a standard passive microrheology test, the movement of dozens of tracers is tracked in a single video frame. The motivation is to average the movements of the tracers and calculate a robust MSD profile.

Observing the MSD for a wide range of integration time scales (or frequencies) gives information on the microstructure of the medium where are diffusing the tracers.

If the tracers are experiencing free diffusion in a purely viscous material, the MSD should grow linearly with sampling integration time:

.

If the tracers are moving in a spring-like fashion within a purely elastic material, the MSD should have no time dependence:

In most cases the tracers are presenting a sub-linear integration-time dependence, indicating the medium has intermediate viscoelastic properties. Of course, the slope changes in different time scales, as the nature of the response from the material is frequency dependent.

Microrheology is another way to do linear rheology. Since the force involved is very weak (order of 10−15 N), microrheology is guaranteed to be in the so-called linear region of the strain/stress relationship. It is also able to measure very small volumes (biological cell).

Given the complex viscoelastic modulus with G(ω) the elastic (conservative) part and G(ω) the viscous (dissipative) part and ω=2πf the pulsation. The GSER is as follows:

with

: Laplace transform of G
kB: Boltzmann constant
T: temperature in kelvins
s: the Laplace frequency
a: the radius of the tracer
: the Laplace transform of the mean squared displacement

A related method of passive microrheology involves the tracking positions of a particle at high frequency, often with a quadrant photodiode. [6] From the position, , the power spectrum, can be found, and then related to the real and imaginary parts of the response function, . [7] The response function leads directly to a calculation of the complex shear modulus, via:

Two Point Microrheology [8]

There could be many artifacts that change the values measured by the passive microrheology tests, resulting in a disagreement between microrheology and normal rheology. These artifacts include tracer-matrix interactions, tracer-matrix size mismatch and more.

A different microrheological approach studies the cross-correlation of two tracers in the same sample. In practice, instead of measuring the MSD , movements of two distinct particles are measured - . Calculating the G(ω) of the medium between the tracers follows:

Notice this equation does not depend on a, but instead in depends on R - the distance between the tracers (assuming R>>a).

Some studies has shown that this method is better in coming to agreement with standard rheological measurements (in the relevant frequencies and materials)

Active microrheology

Active microrheology may use a magnetic field , [9] [10] [11] [8] [12] [13] optical tweezers [14] [15] [16] [17] [18] or an atomic force microscope [19] to apply a force on the tracer and then find the stress/strain relation.

The force applied is a sinusoidal force with amplitude A and frequency ω -

The response of the tracer is a factor of the matrix visco-elastic nature. If a matrix is totally elastic (a solid), the response to the acting force should be immediate and the tracers should be observed moving by-

.

with .

On the other hand, if the matrix is totally viscous (a liquid), there should be a phase shift of between the strain and the stress -

in reality, as most materials are visco-elastic, the phase shift observed is .

When φ>45 the matrix is considered mostly in its "viscous domain" and when φ<45 the matrix is considered mostly in its "elastic domain".

Given a measured response phase shift φ (sometimes noted as δ), this ratio applies:

Similar response phase analysis is used in regular rheology testing.

More recently, it has been developed into Force spectrum microscopy to measure contributions of random active motor proteins to diffusive motion in the cytoskeleton. [4]

Related Research Articles

Quantum teleportation is a technique for transferring quantum information from a sender at one location to a receiver some distance away. While teleportation is commonly portrayed in science fiction as a means to transfer physical objects from one location to the next, quantum teleportation only transfers quantum information. The sender does not have to know the particular quantum state being transferred. Moreover, the location of the recipient can be unknown, but classical information needs to be sent from sender to receiver to complete the teleportation. Because classical information needs to be sent, teleportation can not occur faster than the speed of light.

Uncertainty principle Foundational principle in quantum physics

In quantum mechanics, the uncertainty principle is any of a variety of mathematical inequalities asserting a fundamental limit to the accuracy with which the values for certain pairs of physical quantities of a particle, such as position, x, and momentum, p, can be predicted from initial conditions.

In physics, specifically in quantum mechanics, a coherent state is the specific quantum state of the quantum harmonic oscillator, often described as a state which has dynamics most closely resembling the oscillatory behavior of a classical harmonic oscillator. It was the first example of quantum dynamics when Erwin Schrödinger derived it in 1926, while searching for solutions of the Schrödinger equation that satisfy the correspondence principle. The quantum harmonic oscillator arise in the quantum theory of a wide range of physical systems. For instance, a coherent state describes the oscillating motion of a particle confined in a quadratic potential well. The coherent state describes a state in a system for which the ground-state wavepacket is displaced from the origin of the system. This state can be related to classical solutions by a particle oscillating with an amplitude equivalent to the displacement.

The fluctuation–dissipation theorem (FDT) or fluctuation–dissipation relation (FDR) is a powerful tool in statistical physics for predicting the behavior of systems that obey detailed balance. Given that a system obeys detailed balance, the theorem is a general proof that thermodynamic fluctuations in a physical variable predict the response quantified by the admittance or impedance of the same physical variable, and vice versa. The fluctuation–dissipation theorem applies both to classical and quantum mechanical systems.

Squeezed coherent state type of quantum state

In physics, a squeezed coherent state is a quantum state that is usually described by two non-commuting observables having continuous spectra of eigenvalues. Examples are position and momentum of a particle, and the (dimension-less) electric field in the amplitude and in the mode of a light wave. The product of the standard deviations of two such operators obeys the uncertainty principle:

In physics, the Hanbury Brown and Twiss (HBT) effect is any of a variety of correlation and anti-correlation effects in the intensities received by two detectors from a beam of particles. HBT effects can generally be attributed to the wave–particle duality of the beam, and the results of a given experiment depend on whether the beam is composed of fermions or bosons. Devices which use the effect are commonly called intensity interferometers and were originally used in astronomy, although they are also heavily used in the field of quantum optics.

Jaynes–Cummings model Model in quantum optics

The Jaynes–Cummings model is a theoretical model in quantum optics. It describes the system of a two-level atom interacting with a quantized mode of an optical cavity, with or without the presence of light. It was originally developed to study the interaction of atoms with the quantized electromagnetic field in order to investigate the phenomena of spontaneous emission and absorption of photons in a cavity.

In spectroscopy, the Autler–Townes effect, is a dynamical Stark effect corresponding to the case when an oscillating electric field is tuned in resonance to the transition frequency of a given spectral line, and resulting in a change of the shape of the absorption/emission spectra of that spectral line. The AC Stark effect was discovered in 1955 by American physicists Stanley Autler and Charles Townes.

In physics, a quantum amplifier is an amplifier that uses quantum mechanical methods to amplify a signal; examples include the active elements of lasers and optical amplifiers.

The Kapitza–Dirac effect is a quantum mechanical effect consisting of the diffraction of matter by a standing wave of light. The effect was first predicted as the diffraction of electrons from a standing wave of light by Paul Dirac and Pyotr Kapitsa in 1933. The effect relies on the wave–particle duality of matter as stated by the de Broglie hypothesis in 1924.

Diffusing-wave spectroscopy (DWS) is an optical technique derived from dynamic light scattering (DLS) that studies the dynamics of scattered light in the limit of strong multiple scattering. It has been widely used in the past to study colloidal suspensions, emulsions, foams, gels, biological media and other forms of soft matter. If carefully calibrated, DWS allows the quantitative measurement of microscopic motion in a soft material, from which the rheological properties of the complex medium can be extracted via the microrheology approach.

The SP formula for the dephasing rate of a particle that moves in a fluctuating environment unifies various results that have been obtained, notably in condensed matter physics, with regard to the motion of electrons in a metal. The general case requires to take into account not only the temporal correlations but also the spatial correlations of the environmental fluctuations. These can be characterized by the spectral form factor , while the motion of the particle is characterized by its power spectrum . Consequently, at finite temperature the expression for the dephasing rate takes the following form that involves S and P functions:

In physics, Berry connection and Berry curvature are related concepts which can be viewed, respectively, as a local gauge potential and gauge field associated with the Berry phase or geometric phase. These concepts were introduced by Michael Berry in a paper published in 1984 emphasizing how geometric phases provide a powerful unifying concept in several branches of classical and quantum physics.

The projector augmented wave method (PAW) is a technique used in ab initio electronic structure calculations. It is a generalization of the pseudopotential and linear augmented-plane-wave methods, and allows for density functional theory calculations to be performed with greater computational efficiency.

The semiconductor luminescence equations (SLEs) describe luminescence of semiconductors resulting from spontaneous recombination of electronic excitations, producing a flux of spontaneously emitted light. This description established the first step toward semiconductor quantum optics because the SLEs simultaneously includes the quantized light–matter interaction and the Coulomb-interaction coupling among electronic excitations within a semiconductor. The SLEs are one of the most accurate methods to describe light emission in semiconductors and they are suited for a systematic modeling of semiconductor emission ranging from excitonic luminescence to lasers.

Force Spectrum Microscopy (FSM) is an application of active microrheology developed to measure aggregate random forces in the cytoplasm. Large, inert flow tracers are injected into live cells and become lodged inside the cytoskeletal mesh, wherein it is oscillated by repercussions from active motor proteins. The magnitude of these random forces can be inferred from the frequency of oscillation of tracer particles. Tracking the fluctuations of tracer particles using optical microscopy can isolate the contribution of active random forces to intracellular molecular transport from that of Brownian motion.

In quantum probability, the Belavkin equation, also known as Belavkin-Schrödinger equation, quantum filtering equation, stochastic master equation, is a quantum stochastic differential equation describing the dynamics of a quantum system undergoing observation in continuous time. It was derived and henceforth studied by Viacheslav Belavkin in 1988.

Phonovoltaic

A phonovoltaic (pV) cell converts vibrational (phonons) energy into a direct current much like the photovoltaic effect in a photovoltaic (PV) cell converts light (photon) into power. That is, it uses a p-n junction to separate the electrons and holes generated as valence electrons absorb optical phonons more energetic than the band gap, and then collects them in the metallic contacts for use in a circuit. The pV cell is an application of heat transfer physics and competes with other thermal energy harvesting devices like the thermoelectric generator.

In solid state physics, the Luttinger–Ward functional, proposed by Joaquin Mazdak Luttinger and John Clive Ward in 1960, is a scalar functional of the bare electron-electron interaction and the renormalized many-body Green's function. In terms of Feynman diagrams, the Luttinger–Ward functional is the sum of all closed, bold, two-particle irreducible diagrams, i.e., all diagrams without particles going in or out that do not fall apart if one removes two propagator lines. It is usually written as or , where is the Green's function and is the bare interaction.

The continuous spontaneous localization (CSL) model is a spontaneous collapse model in quantum mechanics, proposed in 1989 by Philip Pearle. and finalized in 1990 Gian Carlo Ghirardi, Philip Pearle and Alberto Rimini.

References

  1. Mason, Thomas G. & Weitz, David A. (1995). "Optical Measurements of Frequency-Dependent Linear Viscoelastic Moduli of Complex Fluids". Physical Review Letters . 74 (7): 1250–1253. Bibcode:1995PhRvL..74.1250M. doi:10.1103/physrevlett.74.1250. PMID   10058972.
  2. Crocker, John C.; Valentine, M. T.; Weeks, Eric R.; Gisler, T.; et al. (2000). "Two-Point Microrheology of Inhomogeneous Soft Materials". Physical Review Letters . 85 (4): 888–891. Bibcode:2000PhRvL..85..888C. doi:10.1103/PhysRevLett.85.888. PMID   10991424.
  3. Levine, Alex J. & Lubensky, T. C. (2000). "One- and Two-Particle Microrheology". Physical Review Letters . 85 (8): 1774–1777. arXiv: cond-mat/0004103 . Bibcode:2000PhRvL..85.1774L. doi:10.1103/PhysRevLett.85.1774. PMID   10970611. S2CID   29305834.
  4. 1 2 Guo, Ming; et al. (2014). "Probing the Stochastic, Motor-Driven Properties of the Cytoplasm Using Force Spectrum Microscopy". Cell. 158 (4): 822–832. doi:10.1016/j.cell.2014.06.051. PMC   4183065 . PMID   25126787.
  5. Furst, Eric M.; Todd M. Squires (2017). Microrheology (First ed.). Oxford. ISBN   0-19-965520-0. OCLC   990115841.
  6. Schnurr, B.; Gittes, F.; MacKintosh, F. C. & Schmidt, C. F. (1997). "Determining Microscopic Viscoelasticity in Flexible and Semiflexible Polymer Networks from Thermal Fluctuations". Macromolecules. 30 (25): 7781–7792. arXiv: cond-mat/9709231 . Bibcode:1997MaMol..30.7781S. doi:10.1021/ma970555n. S2CID   16785228.
  7. Gittes, F.; Schnurr, B.; Olmsted, P. D.; MacKintosh, F. C.; et al. (1997). "Determining Microscopic Viscoelasticity in Flexible and Semiflexible Polymer Networks from Thermal Fluctuations". Physical Review Letters . 79 (17): 3286–3289. arXiv: cond-mat/9709228 . Bibcode:1997PhRvL..79.3286G. doi:10.1103/PhysRevLett.79.3286. S2CID   49563934.
  8. 1 2 Weitz, D. A., John C. Crocker (2000). "Two-Point Microrheology of Inhomogeneous Soft Materials". Phys. Rev. Lett. 85 (4): 888–891. Bibcode:2000PhRvL..85..888C. doi:10.1103/PhysRevLett.85.888. PMID   10991424.
  9. A.R. Bausch; et al. (1999). "Measurement of local viscoelasticity and forces in living cells by magnetic tweezers". Biophysical Journal . 76 (1 Pt 1): 573–9. Bibcode:1999BpJ....76..573B. doi:10.1016/S0006-3495(99)77225-5. PMC   1302547 . PMID   9876170.
  10. K.S. Zaner & P.A. Valberg (1989). "Viscoelasticity of F-actin measured with magnetic microparticles". Journal of Cell Biology . 109 (5): 2233–43. doi:10.1083/jcb.109.5.2233. PMC   2115855 . PMID   2808527.
  11. F.Ziemann; J. Radler & E. Sackmann (1994). "Local measurements of viscoelastic moduli of entangled actin networks using an oscillating magnetic bead micro-rheometer". Biophysical Journal . 66 (6): 2210–6. Bibcode:1994BpJ....66.2210Z. doi:10.1016/S0006-3495(94)81017-3. PMC   1275947 . PMID   8075354.
  12. F. Amblard; et al. (1996). "Subdiffusion and Anomalous Local Viscoelasticity in Actin Networks". Physical Review Letters . 77 (21): 4470–4473. Bibcode:1996PhRvL..77.4470A. doi:10.1103/PhysRevLett.77.4470. PMID   10062546.
  13. Manlio Tassieri; et al. (2010). "Analysis of the linear viscoelasticity of polyelectrolytes by magnetic microrheometry—Pulsed creep experiments and the one particle response". Journal of Rheology . 54 (1): 117–131. Bibcode:2010JRheo..54..117T. doi:10.1122/1.3266946.
  14. E. Helfer; et al. (2000). "Microrheology of Biopolymer-Membrane Complexes" (PDF). Physical Review Letters . 85 (2): 457–60. Bibcode:2000PhRvL..85..457H. doi:10.1103/PhysRevLett.85.457. PMID   10991307.
  15. Manlio Tassieri; et al. (2009). "Measuring storage and loss moduli using optical tweezers: Broadband microrheology". Phys. Rev. E . 81 (2): 026308. arXiv: 0910.1339 . Bibcode:2010PhRvE..81b6308T. doi:10.1103/PhysRevE.81.026308. PMID   20365652. S2CID   9425957.
  16. Daryl Preece; et al. (2011). "Optical tweezers: wideband microrheology". Journal of Optics . 13 (11): 044022. arXiv: 1005.1401 . Bibcode:2011JOpt...13d4022P. doi:10.1088/2040-8978/13/4/044022. S2CID   119286693.
  17. Manlio Tassieri; et al. (2012). "Microrheology with optical tweezers: data analysis". New Journal of Physics . 14 (11): 115032. Bibcode:2012NJPh...14k5032T. doi: 10.1088/1367-2630/14/11/115032 .
  18. David Engström; Michael C.M. Varney; Martin Persson; Rahul P. Trivedi; et al. (2012). "Unconventional structure-assisted optical manipulation of high-index nanowires in liquid crystals". Optics Express . 20 (7): 7741–7748. Bibcode:2012OExpr..20.7741E. doi: 10.1364/OE.20.007741 . PMID   22453452. S2CID   13975852.
  19. Rigato, Annafrancesca; Miyagi, Atsushi; Scheuring, Simon; Rico, Felix (2017-05-01). "High-frequency microrheology reveals cytoskeleton dynamics in living cells". Nature Physics. advance online publication (8): 771–775. Bibcode:2017NatPh..13..771R. doi:10.1038/nphys4104. ISSN   1745-2481. PMC   5540170 . PMID   28781604.