Skip to main content
Log in

Near-wall hindered Brownian diffusion of nanoparticles examined by three-dimensional ratiometric total internal reflection fluorescence microscopy (3-D R-TIRFM)

  • Original
  • Published:
Experiments in Fluids Aims and scope Submit manuscript

Abstract

A three-dimensional nanoparticle tracking technique using ratiometric total internal reflection fluorescence microscopy (R-TIRFM) is presented to experimentally examine the classic theory on the near-wall hindered Brownian diffusive motion. An evanescent wave field from the total internal reflection of a 488-nm bandwidth argon-ion laser is used to provide a thin illumination field on the order of a few hundred nanometers from the wall. Fluorescence-coated polystyrene spheres of 200±20 nm diameter (specific gravity=1.05) are used as tracers and a novel ratiometric analysis of their images allows the determination of fully three-dimensional particle locations and velocities. The experimental results show good agreement with the lateral hindrance theory, but show discrepancies from the normal hindrance theory. It is conjectured that the discrepancies can be attributed to the additional hindering effects, including electrostatic and electro-osmotic interactions between the negatively charged tracer particles and the glass surface.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2a
Fig. 3a, b
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8a, b
Fig. 9a–d

Similar content being viewed by others

Notes

  1. Hosoda et al. (1998) have shown that the near-wall Brownian motion is found to be anisotropic with respect to the directions parallel and perpendicular to the interface using evanescent wave microscopy. Spectroscopic analysis was conducted, allowing a wide range of wave number, by varying the incident ray angle, and the resulting autocorrelation function of the image intensity showed evidence of the anisotropy. However, the scope of their work is far from being comprehensive in that no quantitative measurements of the near-wall hindered diffusion motion have been conducted and compared with the existing theories.

  2. Refer to Fig. 10 in the Appendix for the calculation uncertainties in determining the penetration depth.

  3. Model UP-1830 UNIQ with 1024×1024-pixel CCD elements and each pixel element is of dimension of 6.45×6.45 μm. The camera operates at 30 frames per second, with minimum illumination of 0.04 lux and a signal-to-noise ratio better than 58 dB. A certain level of image smearing because of the finite exposure time is inevitable, and this may result in the blurring of the particle image to a certain degree. However, the finite exposure time does not affect the ratiometric measurements where only the intensity ratios are analyzed and the intensity ratio is unaffected by the image blur. Note that the measured particle location is referred to its closest pole from the solid surface, i.e., the brightest point (refer to Sect. 2.3).

  4. The dye particles are believed to be free from the “photo-bleaching” effect that can render the dye unable to fluoresce after being excessively exposed to high-intensity pumping light. As per the specifications of Molecular Probes (2004), the aqueous suspension of fluorescent beads do not fade noticeably when illuminated by an intense 250-watt xenon-arc lamp for 30 min. Since the current experiment uses approximately 40-mW illumination at 488-nm bandwidth from the 200-mW nominal laser for the total exposure time of up to 4 s, any errors associated with photo-bleaching should be negligibly minimal.

  5. The high-NA objective-based TIRFM system adjusts the incident angle using a fiber optic laser guide attached to a precision positioning system traveling along the barrel axis (http://www.olympusmicro.com/primer/java/tirf/tirfalign/index.html).

  6. The physical location of the brightest particle may not be exactly at zero at the solid surface; rather, it should be at the most probable separation distance. Since both the glass surface and the particles are negatively charged, there exists the most probable separation distance, which is equivalent to the minimum potential energy state ensuring the mechanical equilibrium. Thus, z=0 here indicates just a reference point for the relative locations of other less bright particles.

  7. For dp=200 nm, k=1.38054×10−23 J/K, μ=0.001N s/m for water at T=293 K, the free Brownian diffusivity (Eq. 8) is given as D=2.1451 μm2/s and the averaged square displacement 〈Δz2〉=2DΔt=0.142 μm2 for the time interval of 33 ms of the 30 fps imaging. The average displacement 〈|Δz|〉 can be approximated to \({\sqrt {{\left\langle {\Delta z^{2} } \right\rangle }} } = 377\;{\text{nm}}{\text{.}} \) The near-wall hindered displacement (Fig. 5) will reduce it to about a half, i.e., 〈Δz〉~±189 nm, which occupies approximately 70% of zp.

  8. When the ray angle increases to 65°, the uncertainty is noticeably reduced to ±6.85 nm, but its penetration depth will not be sufficient to accommodate the pertinent Brownian motion length scale.

  9. Meiners and Quake (1999) attempted a direct measurement of hydrodynamic interaction between two spherical colloid particles, ranging from 3.1 µm to 9.8 µm, two order of magnitudes larger than the present nanoparticles, suspended by optical tweezers in an external potential. However, they measured the cross-correlations of only two-dimensional motions of particles without accounting for the near-wall hindrance.

References

  • Axelrod D, Burghardt TP, Thompson NL (1984) Total internal reflection fluorescence (in biophysics). Annu Rev Biophys Bio 13:247–268

    CAS  Google Scholar 

  • Axelrod D, Hellen EH, Fulbright RM (1992) Total internal reflection fluorescence. In: Lakowicz J (ed) Topics in fluorescence spectroscopy: principles and applications, vol 3: biochemical applications. Plenum Press, New York, pp 289–343

  • Banerjee A, Chon C, Kihm KD (2003) Nanoparticle tracking using TIRFM imaging. In: Photogallery of the ASME international mechanical engineering congress and exposition (IMECE 2003), Washington, DC, November 2003

  • Batchelor GK (1975) Brownian diffusion of particles with hydrodynamic interaction. J Fluid Mech 74:1-29

    Google Scholar 

  • Bevan MA, Prieve DC (2000) Hindered diffusion of colloidal particles very near to a wall: revisited. J Chem Phys 113(3):1228–1236

    Article  CAS  Google Scholar 

  • Born M, Wolf E (1980) Principles of optics, 6th edn. Cambridge University Press, Cambridge, pp 47–51

  • Brenner H (1961) The slow motion of a sphere through a viscous fluid towards a plane surface. Chem Eng Sci 16:242–251

    Article  CAS  Google Scholar 

  • Brown R (1828) A brief account of microscopical observations made in the months of June, July, and August, 1827, on the particles contained in the pollen of plants; and on the general existence of active molecules in organic and inorganic bodies. Phil Mag 4:161–173

    Google Scholar 

  • Burghart TP, Thompson NL (1984) Effects of planar dielectric interfaces on fluorescence emission and detection. Biophys J 46:729–737

    PubMed  Google Scholar 

  • Deen WM (1998) Analysis of transport phenomena. Oxford University Press, New York, pp 59–63

  • Denk W, Strickler JH, Webb WW (1990) Two-photon laser scanning fluorescence microscopy. Science 248:73–76

    CAS  PubMed  Google Scholar 

  • Dufresne ER, Squires TM, Brenner MP, Grier DG (2000) Hydrodynamic coupling of two Brownian spheres to a planar surface. Phys Rev Lett 85:3317–3320

    Article  CAS  PubMed  Google Scholar 

  • Einstein A (1905) Über die von der molekularkinetischen Theorie der Wärme geforderte Bewegung von in ruhenden Flüssigkeiten suspendierten Teilchen. Ann Physik 17: 549

    CAS  Google Scholar 

  • Einstein A (1956) Investigations on the theory of Brownian movement. Dover, New York

  • Fox RW, McDonald AT, Pritchard PJ (2004) Introduction to fluid mechanics, 6th edn. Wiley, Hoboken, New Jersey, pp 755–761

  • Goldman AJ, Cox RG, Brenner H (1967) Slow viscous motion of a sphere parallel to a plane. I: motion through a quiescent fluid. Chem Eng Sci 22:637–651

    Article  CAS  Google Scholar 

  • Goos VF, Hanchen H (1947) Ein neuer und fundamentaler versuch zur totalreflexion. Ann Physik 1:333–346

    Google Scholar 

  • Hecht E (2002) Optics, 4th edn. Addison-Wesley, Reading, Massachusetts, pp 124–127

  • Hellen EH, Axelrod D (1986) Fluorescence emission at dielectric and metal film interfaces. J Opt Soc Am B 4:337–350

    Google Scholar 

  • Hosoda M, Sakai K, Takagi K (1998) Measurement of anisotropic Brownian motion near an interface by evanescent light-scattering spectroscopy. Phys Rev E 58(5):6275–6280

    Article  CAS  Google Scholar 

  • Ingenhousz J (1779) Experiments on vegetables, discovering their great power of purifying the common air in sunshine, and of injuring it in the shade or at night. In: Marshall HL, Herbert SK (1952) A source book in chemistry: 1400–1900. Harvard University Press, Cambridge, Massachusetts

    Google Scholar 

  • Inoue S (1987) Video microscopy. Plenum Press, New York

  • Ishijima A, Yanagida T (2001) Single molecule nanobioscience. Trends in Biochem Sci 26:438–444

    Article  CAS  Google Scholar 

  • Jin S, Huang P, Park J, Yoo JY, Breuer KS (2003) Near-surface velocimetry using evanescent wave illumination. In: Proceedings of the ASME international mechanical engineering congress and exposition (IMECE 2003), Washington, DC, November 2003, paper 44015

  • Kim MJ, Beskok A, Kihm KD (2002) Electro-osmosis-driven micro-channel flows: a comparative study of microscopic particle image velocimetry measurements and numerical simulations. Exp Fluids 33:170–180

    CAS  Google Scholar 

  • Kim S, Karrila SJ (1991) Microhydrodynamics: principles and selected applications. Butterworth-Heinemann, Stoineham, Massachusetts

  • Kline SJ, McClintock FA (1953) Describing uncertainties in single-sample experiments. Mech Eng 75(1): 3–9

    Google Scholar 

  • Kohonen T (1995) Self-organizing maps. Springer, Berlin Heidelberg, New York

  • Kohonen T, Kaski S, Lappalainen H (1994) Self-organized formation of various invariant-feature filters in the adaptive subspace SOM. Neural Comput 9:1321–44

    Google Scholar 

  • de Lange F, Cambi A, Huijbens R, de Bakker B, Rensen W, Garcia-Parajo M, van Hulst N, Figor CG (2001) Cell biology beyond the diffraction limit: near-field scanning optical microscopy. J Cell Sci 114:4153–4160

    PubMed  Google Scholar 

  • Meiners J-C, Quake SR (1999) Direct measurement of hydrodynamic cross correlations between two particles in an external potential. Phys Rev Lett 82(10):2211–2214

    Article  CAS  Google Scholar 

  • Molecular Probes (2004) Personal communications, Molecular Probes Inc., Eugene, Oregon

  • Nakroshis P, Amoroso M, Legere J, Smith C (2003) Measuring Boltzmann’s constant using video microscopy of Brownian motion. Am J Phys 71(6):568–573

    Article  CAS  Google Scholar 

  • Okamoto K, Hassan YA, Schmidl WD (1995) New tracking algorithm for particle image velocimetry. Exp Fluids 19:342–347

    Article  Google Scholar 

  • Okamoto K, Nishio S, Kobayashi T, Saga T (1997) Standard images for particle imaging velocimetry. In: Proceedings of the 2nd international workshop on particle image velocimetry (PIV’97), Fukui, Japan, July 1997, pp 229–236

  • Park JW, Choi CK, Kihm KD (2004) Optically slices micro-PIV using confocal laser scanning microscopy (CLSM). Exp Fluids 37:105–119

    Article  Google Scholar 

  • Pawley JB (1995) Handbook of biological confocal microscopy, 2nd edn. Plenum Press, New York

  • Prieve DC (1999) Measurement of colloidal forces with TIRM. Adv Colloid Interfac 82:93–125

    Article  CAS  Google Scholar 

  • Probstein RF (1994) Physicochemical hydrodynamics. Wiley, New York

  • Rohrbach A (2000) Observing secretory granules with a multiangle evanescent wave microscope. Biophys J 78:2641–2654

    CAS  PubMed  Google Scholar 

  • Sako Y, Minoghchi S, Uyemura T (2000) Single-molecule imaging of EGFR signaling on the surface of living cells. Nat Cell Biol 2:168–172

    Article  CAS  PubMed  Google Scholar 

  • Sako Y, Yanagida T (2003) Single-molecule visualization in cell biology. Nat Rev Mol Cell Bio September supplement SS1-SS5

  • Salmon R, Robbins C, Forinash K (2002) Brownian motion using video capture. Eur J Phys 23(3):235-253

    Article  Google Scholar 

  • Schatzel K, Neumann WG, Muller J, Materzok B (1992) Optical tracking of Brownian particles. Appl Optics 31:770–778

    Google Scholar 

  • Shlesinger MF, Klafter J, Zumofen G (1999) Above, below and beyond Brownian motion. Am J Phys 67:1253–1259

    Article  Google Scholar 

  • Stelzer EHK, Lindek S (1994) Fundamental reduction of the observation volume in far-field light microscopy by detection orthogonal to the illumination axis: confocal theta microscopy. Opt Commun 111:536–547

    Article  Google Scholar 

  • Takagi T, Okamoto K (2001) Particle tracking velocimetry by network model. In: Proceedings of the 3rd Pacific symposium on flow visualization and image processing (PSFVIP-3), Maui, Hawaii, March, conference CD-ROM

  • Weiss S (2000) Measuring conformational dynamics of biomolecules by single molecule fluorescence spectroscopy. Nat Struct Biol 7:724–729

    Article  CAS  PubMed  Google Scholar 

  • Xie S (2001) Single-molecule approach to enzymology. Single Mol 2:229–236

    Article  CAS  Google Scholar 

  • Zettner CM, Yoda M (2003) Particle velocity field measurements in a near-wall flow using evanescent wave illumination. Exp Fluids 34:115–121

    CAS  Google Scholar 

Download references

Acknowledgment

The authors wish to thank Mr. Eiji Yokoi of Olympus America Inc. for his technical assistance in setting up the TIRFM system. The authors are grateful to the financial support sponsored partially by the NASA-Fluid Physics Research Program grant no. NAG 3–2712, and partially by the US-DOE/Argonne National Laboratory grant no. DE-FG02–04ER46101. The presented technical contents are not necessarily the representative views of NASA, US-DOE, or the Argonne National Laboratory.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to K. D. Kihm.

Uncertainty analyses

Uncertainty analyses

Experimental uncertainties have been estimated based on a single-sample experiment, where only one measurement is made for each point (Kline and McClintock 1953). Four pertinent uncertainties are presented: (1) uncertainty for incident ray angle θ; (2) uncertainty for the lateral (xy) Brownian displacement measurements Δx or Δy; (3) uncertainty for the penetration depth zp; and (4) uncertainty for the z directional relative displacement Δh.

1.1 Uncertainty for incident ray angle

The evanescent wave rim radius at the back-focal plane of the lens, R (http://www.olympusmicro.com), is given as R=fnsinθ, where the focal length of the TIRF objective lens f=3 mm and the refractive index of the incident cover glass medium n=1.515. A measurement function for the incident ray angle is defined as sinθ=R/fn=g(R, f, n). The Kline-McClintock analysis (Fox et al. 2004) for uncertainty w g gives:

$$\begin{array}{*{20}l} {{w_{g} } \hfill} & {{ = \pm {\left[ {{\left( {\frac{{\partial g}}{{\partial R}}w_{R} } \right)}^{2} + {\left( {\frac{{\partial g}}{{\partial f}}w_{f} } \right)}^{2} + {\left( {\frac{{\partial g}}{{\partial n}}w_{n} } \right)}^{2} } \right]}^{{\frac{1}{2}}} } \hfill} \\ {{} \hfill} & {{ = \pm {\left[ {{\left( {\frac{{w_{R} }}{R}g} \right)}^{2} + {\left( { - \frac{{w_{f} }}{f}g} \right)}^{2} + {\left( { - \frac{{w_{n} }}{n}g} \right)}^{2} } \right]}^{{\frac{1}{2}}} } \hfill} \\ {{} \hfill} & {{ = \pm {\left[ {{\left( {\frac{{w_{R} }}{{fn}}} \right)}^{2} + {\left( {\frac{{w_{f} \sin \theta }}{f}} \right)}^{2} + {\left( {\frac{{w_{n} \sin \theta }}{n}} \right)}^{2} } \right]}^{{\frac{1}{2}}} } \hfill} \\ \end{array} $$
(18)

where w R , w f , and w n are the uncertainties associated with the individual parameters of R, f, and n. Per the resolution limit provided by Olympus Inc., w R =±0.025 mm, and both the focal length uncertainty, w f , and the refractive index uncertainty, w n , are assumed to be negligibly small. For the incident angle of 62°, the resulting uncertainty is calculated as \(w_{\theta } = \pm 0.315^\circ \) after a conversion based on w θ =sin−1(w g ).

1.2 Uncertainty for the lateral (xy) Brownian displacement measurements

A reasonable estimate of the measurement uncertainty due to random error is plus or minus half of the smallest scale division, equivalent to 1 pixel, of the CCD camera. By taking the average pixel displacements as 3 pixel, the lateral displacement uncertainty is estimated to be \( w_{x} = w_{y} = \pm 0.5\;{\text{pixel}}\; \cong \; \pm 71.7\;{\text{nm}}. \)

1.3 Uncertainty for the penetration depth zp

Considering Eq. 2:

$$z_{{\text{p}}} = \frac{{\lambda _{0} }}{{4\pi }}{\left( {n^{2}_{i} \sin ^{2} \theta - n^{2}_{t} } \right)}^{{ - \frac{1}{2}}} $$

the uncertainty equation for zp is given as:

$$w_{{z_{{\text{p}}} }} = \pm {\left[ {{\left( {\frac{{\partial z_{{\text{p}}} }}{{\partial \lambda _{0} }}w_{{\lambda _{0} }} } \right)}^{2} + {\left( {\frac{{\partial z_{{\text{p}}} }}{{\partial n_{i} }}w_{{n_{i} }} } \right)}^{2} + {\left( {\frac{{\partial z_{{\text{p}}} }}{{\partial n_{t} }}w_{{n_{t} }} } \right)}^{2} + {\left( {\frac{{\partial z_{{\text{p}}} }}{{\partial \theta }}w_{\theta } } \right)}^{2} } \right]}^{{\frac{1}{2}}} $$
(19)

where the optical blue filter for the laser beam has a bandwidth of \(w_{{\lambda _{0} }} = \pm 2\;{\text{nm}},\;w_{\theta } = \pm 0.315^\circ \) (first section of Appendix), and variations of refractive indices are neglected, i.e., \(w_{{n_{i} }} = w_{{n_{t} }} = 0. \) The measurement uncertainty of zp shows a significant increase with the incident ray angle θ approaching the critical value of θc=61.38° (Fig. 10). For example, the penetration depth uncertainty is estimated as \(w_{{z_{{\text{p}}} }} = \pm 4.75\;{\text{nm}} \) for θ=65°, but increases to \(w_{{z_{{\text{p}}} }} = \pm 69.47\;{\text{nm}} \) for θ=62°.

Fig. 10
figure 10

Calculated uncertainty for the penetration depth (zp) for different incident angles (θ) with θc=61.38° for the glass–water interface (n i =1.515 for glass, n t =1.33 for water)

1.4 Uncertainty for the line-of-sight (z) Brownian displacement measurements

The uncertainty of Δh can be estimated by applying the uncertainty analysis to the ratiometric intensity relation given in Eq. 6 as:

$$RaInt = \frac{{I^{1}_{N} {\left( {h_{1} ,\;R,\;c} \right)}}}{{I^{2}_{N} {\left( {h_{2} ,\;R,\;c} \right)}}} = {\text{e}}^{{{\left( { - \frac{{\Delta h}}{{z_{{\text{p}}} }}} \right)}}} $$

or equivalently:

$$\Delta h = - z_{{\text{p}}} \ln {\left( {RaInt} \right)} = z_{{\text{p}}} {\left[ {\ln {\left( {I^{2}_{{\text{N}}} } \right)} - \ln {\left( {I^{1}_{{\text{N}}} } \right)}} \right]} $$
(6a)

Thus, the uncertainty equation is obtained as:

$$w_{{\Delta h}} = \pm {\left[ {{\left( {\frac{{\partial \Delta h}}{{\partial z_{{\text{p}}} }}w_{{z_{{\text{p}}} }} } \right)}^{2} + {\left( {\frac{{\partial \Delta h}}{{\partial I^{2}_{{\text{N}}} }}w_{{I^{2}_{{\text{N}}} }} } \right)}^{2} + {\left( {\frac{{\partial \Delta h}}{{\partial I^{1}_{{\text{N}}} }}w_{{I^{1}_{{\text{N}}} }} } \right)}^{2} } \right]}^{{\frac{1}{2}}} $$
(20)

The elementary uncertainties of \(w_{{I^{1}_{{\text{N}}} }} \;{\text{and}}\;w_{{I^{2}_{{\text{N}}} }} \) may be estimated from the statistical nature of the measured data. A statistical analysis was conducted for the particle tracking data for θ i =62° and the resulting statistical properties are summarized in terms of the pixel gray level, as shown in Table 2. Thus, the elementary uncertainties for the image intensities is assumed to be equal to the width of a 95% confidence interval, i.e., \(w_{{I^{1}_{{\text{N}}} }} = w_{{I^{2}_{{\text{N}}} }} = \pm 10.6835 \) (pixel equivalency), the reference particle image intensity \(I^{2}_{{\text{N}}} \) is assumed to have the maximum intensity of 220, and the arbitrary particle image intensity \(I^{1}_{{\text{N}}} \) is assumed to have the average intensity of 165.94. Substituting these values into Eq. 20, the overall uncertainty for the z location measurements is estimated as \(w_{{\Delta h}} = \pm 29.38\;{\text{nm}} \) for θ=62°. This R-TIRFM uncertainty also decreases with increasing incident ray angle, depicting a similar trend of the penetration depth uncertainty.

Table 2 Statistical properties of the analysis conducted for the particle tracking data for θ i =62°

Rights and permissions

Reprints and permissions

About this article

Cite this article

Kihm, K.D., Banerjee, A., Choi, C.K. et al. Near-wall hindered Brownian diffusion of nanoparticles examined by three-dimensional ratiometric total internal reflection fluorescence microscopy (3-D R-TIRFM). Exp Fluids 37, 811–824 (2004). https://doi.org/10.1007/s00348-004-0865-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00348-004-0865-4

Keywords

Navigation