Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Photoacoustic measurement of refractive index of dye solutions and myoglobin for biosensing applications

Open Access Open Access

Abstract

Current methods of determining the refractive index of chemicals and materials, such as ellipsometry and reflectometry, are limited by their inability to analyze highly absorbing or highly transparent materials, as well as the required prior knowledge of the sample thickness and estimated refractive index. Here, we present a method of determining the refractive index of solutions using the photoacoustic effect. We show that a photoacoustic refractometer can analyze highly absorbing dye samples to within 0.006 refractive index units of a handheld optical refractometer. Further, we use myoglobin, an early non-invasive biomarker for malignant hyperthermia, as a proof of concept that this technique is applicable for use as a medical diagnostic. Comparison of the speed, cost, simplicity, and accuracy of the techniques shows that this photoacoustic method is well-suited for optically complex systems.

© 2013 Optical Society of America

1. Introduction

The refractive index of a material describes the group velocity of light as it propogates through that medium. This property is highly correlated with both the density of the medium, and the absorptive properties of the medium. Using Kramers-Kroenig analysis (mathematical relationships that relate the complex and real parts of a physical system) it is possible to estimate many material properties, including refractive index. However, in recent years, the legitimacy of this analysis has been questioned for heterogenous biological materials, such as hemaglobin, due to the complex nature of the analysis [1]. Measurement of the refractive index is based upon the fundamental Fresnel equations, which describe the refraction, reflection, and transmission of light at an interface between two media with non-matching refractive indices. In recent years, the gold standard of measurement of this property has been either reflectometry or ellipsometry, depending upon various sample properties, such as its transparency, and thickness, as well as overall cost, and the necessary data precision. However, for highly absorbing materials, partially transparent materials, or entirely unknown materials, both techniques demonstrate significant drawbacks that prevent the widespread adoption of refractive index measurements beyond typical materials characterization. Here, we explore a low cost, simple, alternative method based on the photoacoustic effect. We demonstrate that our photoacoustic method, Photoacoustic Spectroscopy/Total Internal Reflection Photoacoustic Spectroscopy (PAS/TIRPAS) refractometry, can be successfully used to determine the refractive index of both dyes, as well as a physiologically relevant concentration of myoglobin [2, 3] in phosphate buffered saline.

1.1. Reflectometry

Reflectometry is a technique in which a polarized beam of light, operating in two linear polarizations, S and P, optically interrogates a sample. The reflected light from the sample is then measured by a photodetector. Both S and P polarizations, along with the incident angle, are recorded in order to obtain the real component of the refractive index. Equation 1 shows how the refractive index of a material is related to the reflection coefficients and angle of incidence.

nR=sinθi[1+(1ρ1+ρ)2tan2θi]12
where ρ=RpRs and nR=n2n1. Equation 1 is used to determine the real part of the refractive index of a material using reflectometry methods [4].

1.2. Critical angle analysis

In a special case, the real component of the refractive index can be determined by measuring the critical angle, or the angle at which total internal reflection occurs. This angle can be determined using equation 7. This method simplifies the analysis of Fresnel’s equations and allows the refractive index of the sample to be found directly, if the refractive index of the substrate, n1, and the angle of incidence, θ1, are known.

1.3. Ellipsometry

Ellipsometry, a closely related technique to reflectometry, can be described with a slight alteration to Equation 1. The relative permittivity of a material, ε̃, can be found, and is directly related to the refractive index of a material by Equation 3 as explained in [6].

ε˜=sin2ϕ[1+(1ρ1+ρ)2tan2ϕ]
ε˜=n2=κ
where ε̃ and κ are the relative permittivity of a material, n is the refractive index of the sample material, ϕ is the incident angle on the sample, and ρ=RpRs is the ratio between the two reflected values for both linear polarizations, P and S. For simple bulk substrate systems, ellipsometry uses the same equation as reflectometry, albeit formatted differently, to determine the refractive index of a material.

1.4. Limitations

Each of these methods has limitations. Reflectometry’s main limitation lies in the fact that the technique does not take into account surface roughness or optical scattering effects. Therefore, the actual measurement of the refractive index is a combination of the optical contributions from both refractive index and surface or substrate roughness. Additionally, reflectometry typically has less success with materials that have many highly absorbing layers, due to the light being unable to be measured from being completely absorbed.

Ellipsometry has similar issues with measuring materials that are highly absorbing for the same reasons. Additionally, ellipsometry has the additional limitation of being unable to determine a single optical property independently of the thickness of a material. This can lead to erroneous results if the operator is not appropriately trained and makes incorrect assumptions or initial guesses in the curve fitting program to determine refractive index and layer thickness.

Critical angle analysis is highly accurate and inexpensive, however, it requires a prism setup which can be optically complicated and physically large. Additionally, the material of the prism must be of a larger refractive index than the sample in order for the condition for total internal reflection to be fulfilled, which poses problems for unknown samples and increases cost.

1.5. The photoacoustic effect

An alternative to all of these methods relies on the photoacoustic effect. The photoacoustic effect is generated by the absorption of light into a sample, which induces pressure in the form of an acoustic wave. The simplest equation to describe photoacoustic excitation is equation 4 from Viator et al. [13] that describes the pressure distribution of a generated photoacoustic plane wave. The magnitude of the resulting acoustic wave relies directly upon the optical absorption coefficient of the material.

p0(z)=μaΓH0eμaz
μa=1δ
where z represents physical depth, μa is the optical absorption coefficient, G is the Grunesian coefficient, H0 is the incident laser energy per area, δ is the optical penetration depth, and p0 is the photoacoustic pressure produced. The inverse of the optical absorption coefficient as shown in equation 5 is the optical penetration depth or measure of the distance the light travels until it decays to 1e of its initial optical intensity. Equation 4 with various experimental techniques are used to analyze the thickness, absorption, and imaging of many materials from blood vessels to thin films. [712]

1.6. Total Internal Reflection Photoacoustic Spectroscopy (TIRPAS)

When the refractive index of one medium is larger than that of a second medium, Snell’s law, equation 6, shows that beyond the critical angle, θc, light is totally internally reflected back into the first medium.

n1sinθ1=n2sinθ2

When θθc, and sinθ2 → 1

θc=sin(n2n1)1

This configuration produces total internal reflection, and generates a special type of nearfield optical effect called the evanescent field. The evanescent field can be perturbed to reduce the reflected coefficient of light if a sample material of appropriate optical absorption or scattering is sufficiently near the field. This type of frustrated total internal reflection is the basis of many characterization systems including attenuated total reflectance [5].

Recently, a method of generating a photoacoustic response from optical absorption through a pulsed evanescent field was rediscovered by Goldschmidt et al. [14, 15], who used the technique in conjunction with modern lasers to improve the method’s sensitivity. This technique is graphically demonstrated in Fig. 1. The method has been used successfully to detect β -hematin to demonstrate the technique’s potential as a new type of biosensor [14].

 figure: Fig. 1

Fig. 1 Schematic of setup showing how laser light interrogates sample to give acoustic effect in TIRPAS

Download Full Size | PDF

In the 1980’s, this technique [1626] was explored as an alternative method to photoacoustic spectroscopy, due to total internal reflection photoacoustic spectroscopy’s (TIRPAS) ability to determine information about materials at an interface. This type of information was never achievable with normal incidence photoacoustic spectroscopy (PAS), due to the generally large optical penetration depth obtained with low absorption materials using the PAS method as can be seen in Fig. 2. TIRPAS was used to determine dye adsorption properties at a glass/solution interface for interfacial phenomena studies as part of a cyclic voltammetry project.

 figure: Fig. 2

Fig. 2 Figure illustrates the vastly different optical absorption depths between the PAS and TIRPAS excitation conditions.

Download Full Size | PDF

The excitation of the photoacoustic effect using TIRPAS is fundamentally different than that by typical PAS experiments due to the fact that the evanescent field is interacting with absorbers close to the interface as opposed to the photoacoustic effect being excited by propagating photons through normal incidence. This type of excitation limits the optical penetration depth to the penetration of the evanescent field which is nearly always much less than the optical penetration depth of propagating photons through a typical material. The evanescent field component in polarization parallel to the optic is given by Equation 8 which comes from Equation 9 and Equation 10, as presented by Muessig et al. [18].

E||=(Ex2+Ez2)12
Ex=2E||(sin2θn212)12cosθ(1n212)12[(1+n21)sin2θn212]12
Ez=2E||sinθcosθ(1n212)12[(1+n21)sin2θn212]12
where E|| is the component of the electric field parallel to the plane of incidence, n21=n2n1, θ is the angle of incidence of the laser on the prism/sample interface, and E′|| is the parallel component of the evanescent field. Aside from the electric field components, which may affect the TIRPAS signal, the evanescent field’s near field interaction can be described by a parameter called the penetration depth, as noted by Goldschmidt et al. [14]. The penetration depth shown in Equation 11 describes the point at which the non-propagating evanescent field decays to 1e of its original intensity. This distance can be roughly used as the depth that can be analyzed with the photoacoustic effect in TIRPAS.
δp=λ12π(sin2θn212)12
where λ1=λn1, λ is the vacuum wavelength of laser used and δp is the optical penetration depth of the evanescent field. This relationship is important because it implies that the signal magnitude in TIRPAS is proportional to the amount of material excited assuming the material is sufficiently thicker than the penetration depth of the evanescent field.

1.7. Surface Plasmon Resonance Photoacoustic Spectroscopy (SPRPAS)

Another method, related to TIRPAS, was explored in the early 1980s that used the resonant effect of surface plasmons to induce a photoacoustic effect for the detection of materials at an interface as well. Several groups [2743] used this approach to understand the technique and to detect various materials, such as nicotine, and various gases of interest. The optical setups varied significantly, however, these methods generally used a laser pulse not meeting the excitation requirement for thermal confinement in the samples of interest. The entire systems were enclosed in photoacoustic cells for the detection and characterization of materials with properties related to and including refractive index. This technique, however, differs from TIRPAS in that TIRPAS uses pulsed laser light which interacts with a sample directly by the evanescent field. The related surface plasmon technique uses the generated evanescent field with an intervening metal layer to generate surface plasmons for detection of materials adsorbed on the surface of the metal layer.

1.8. PAS/TIRPAS refractometry

PAS/TIRPAS refractometry is a new technique that combines the advantages of both the photoacoustic effect and the evanescent field to determine the refractive index of materials. It should be noted that a similar, but more computationally and physically complicated non-thermally confined photoacoustic refractometry technique was accomplished in the early 1990s. In that case, researchers relied upon the analysis of resonant optical modes in Langmuir-Blodgett films [4447] and were able to produce refractive index estimates on the order of 0.01. In traditional reflectometry, light is measured after its interaction with a sample in two linear polarization states to determine refractive index. In PAS/TIRPAS refractometry, a laser beam is scanned across a sample at changing angles of incidence to collect acoustic waves from both the PAS and TIRPAS regimes, as well as any location inbetween. This experiment allows for the direct observation of the light-induced acoustic wave, which changes drastically in peak to peak voltage between the PAS and TIRPAS regimes. This was shown experimentally in previous data from Goldschmidt et al. [14], and can be physically shown to be different by Fig. 2. The drastic change in peak to peak voltage can be easily explained by comparing equations 5 and 4 with equation 11 which for typical values gives an order of magnitude or more difference in photoacoustic pressure.

Here we demonstrate a simple, robust technique based upon the large changes in optical penetration depth for the determination of the real component of the refractive index of dyed liquids and the biologically relevant material, myoglobin, for potential disease detection.

2. Materials and methods

2.1. Materials

2.1.1. Direct Red dye 81 solutions

The dye solutions were made with a 50% polyethylene glycol (PEG, m.w. 200) 50% H2O solution along with the dye at 125, 250, 500, and 750 μg/ml respectively. The PEG was added in order to shift the refractive index closer to the middle range of the ATAGO R-5000 refractometer to provide easier measurement and comparison with the PAS/TIRPAS refractometer. Each solution was mixed for at least 5 minutes before testing using a hot plate/stirrer (Corning PC 220). The mixing directly preceeded each test to reduce the effects from dye settling on the prism surface.

2.1.2. Myoglobin solution

The myoglobin solution was made by mixing 460 μg of dry horse skeletal myoglobin (Sigma Aldrich #M0630-250MG) per 1 ml of phosphate buffered saline (Fisher #BP2438-4). As with the dye solutions, the myoglobin was mixed for at least 5 minutes before testing using a hot plate/stirrer. The myoglobin solution did not have added PEG to it since the refractive index was within the readable scale on the handheld refractometer.

2.2. PAS/TIRPAS refractometer

The device, shown in Fig. 3, works on the principle of exciting a photoacoustic effect in the sample at different angles of incidence. The samples were interrogated with vertically polarized 532 nm laser light generated by a Q-switched Nd:YAG laser (Surelite I-20 Continuum). The acoustic signal was detected using a 15 MHz transducer with a 1” focal length (Olympus #U8472001).

 figure: Fig. 3

Fig. 3 3-dimensional model of the setup

Download Full Size | PDF

2.2.1. Laser leveling

The PAS/TIRPAS holder was leveled relative to the optical table by a Micromark digital level (Micromark #: 84519). The interrogating laser beam was leveled relative to the table as well by a setup similar to a Michaelson interferometer except that the output of the setup was placed on a flourescent card where the two beams could be made coincident to level the beam. Thorlab’s mirrors, a beam splitter, and cage system connectors were used to create this setup (Thorlabs #: CM1-BS013 and #: PF10-03-G01). The PAS/TIRPAS refractometer was aligned by viewing the incident and reflected spots on a sheet of fluorescent paper and adjusting the incident reflected laser to overlay it on that of the reflected beam.

2.2.2. Labview automation

A Sherline stepper motor XY base (Sherline 5200-CNC) was used, along with a 4” rotary table (P/N 8700), right angle attachment (P/N 3701), and 3-jaw chuck (P/N 1041) to hold and position the rotational mount made of a prism and custom machined PTFE rodstock. The rotary table was controlled by Labview. The angular spectrum settings were set to scan a range of angles where the critical angle was initially assumed to be. The Labview program stepped the motor at 0.01 degree increments across the entire range of angles. At each increment, the Labview program measured as stated the photoacoustic wave’s peak to peak voltage.

2.2.3. Data analysis

The angular spectra were analyzed by taking the derivative of the angular spectrum data to find the local minima corresponding to the reduction in photoacoustic signal amplitude. The process of taking the derivative of the angular spectrum data produced fairly noisy graphs that needed further data smoothing to correctly determine the critical angle. Kaleidagraph 4.1.0 was used with the smooth curve fit function to smooth the data so that the critical angle could be extracted from the data. This smooth curve fit function is based upon a combination of Stineman interpolation [48] combined with a weighting structure to smooth the data. The exact formula is described further in the appendix.

2.2.4. PAS/TIRPAS measurement

First, the laser beam was leveled using the technique described in section 2.2.1. Next, the sample holder was filled with sample liquid and the sample holder was leveled using the Micromark digital level by balancing the level on the unconnected acoustic sensor which was previously inserted into the sample holder. Finally, the Labview software described in section 2.2.2 was used along with the Sherline stepper motors to obtain the angular spectrum scan of the sample material. The angular scan was then used to determine the critical angle as explained in section 2.2.3.

2.3. Optical refractometer measurement

We used an ATAGO R-5000 refractometer along with a 532 nm line filter (Thorlabs #: FL532-10) and a polarizing beam splitter cube (Thorlabs #: CM1-PBS251) to limit the polarization and wavelengths measured to find the refractive index of the samples for comparison with our PAS/TIRPAS refractometry technique as shown in Fig. 4.

 figure: Fig. 4

Fig. 4 Refractometer setup used to compare refractive index with PAS/TIRPAS refractometry

Download Full Size | PDF

3. Results

A typical result from the dye is shown in Fig. 5(a) and demonstrates a clear transition in photoacoustic peak to peak voltage at approximately 66 degrees. The derivative of the angular spectrum data was taken along with smoothing to obtain a more precise measurement of the transition as demonstrated by the arrow in Fig. 5(b). Figure 5(c) shows a second smoothing applied to the data in order to obtain the minima corresponding to the critical angle without shifts due to noise. Finally, it is straightforward to determine the critical angle from finding the minima in Fig. 5(d).

 figure: Fig. 5

Fig. 5 Graphs a, b, c, and d show the step by step data analysis process to determine the critical angle.

Download Full Size | PDF

In order to demonstrate the utility of PAS/TIRPAS refractometry, we compared the results obtained with optical reference measurements with the ATAGO R-5000 refractometer setup. We show agreement between the two techniques to 0.006 refractive index units between the samples’ refractive index values using PAS/TIRPAS refractometry and the ATAGO R-5000 refractometer as shown in table 1.

Tables Icon

Table 1. PAS/TIRPAS refractometry as compared to a handheld refractometer demonstrates that our method is in good agreement with standard techniques.

4. Discussion

As was shown in our results section, PAS/TIRPAS refractometry is an accurate and precise alternative to ellipsometry and reflectometry. Every dye solution’s refractive index measurement fell within the measurement range of the handheld refractometer. In the case of the Direct Red, 750 μg/ml, the sample-to-sample variability seemed to be more significant (0.005) than the other dyes. We believe this was due to dye settling over time. The myoglobin concentration seemed to have some divergence from the measured value with our handheld refractometer, however, this variability can be explained by the lower signal-to-noise ratio as shown in Fig. 6. However, variability of 0.006 is still fairly reasonable compared to the differences between reflectometry and ellipsometry for many materials and therefore this is not significant for our current demonstration of the technique’s capabilities. Therefore, PAS/TIRPAS refractometry has the capability to analyze both dyes and a biologically relevant analyte, myoglobin, for potential biosensing and biomedical monitoring. Our technique opens up significant potential for biomedical monitoring and adsorption state monitoring through the refractive index at a sample/prism interface.

 figure: Fig. 6

Fig. 6 Acoustic differences between dye solution and myoglobin solution in the PAS regime

Download Full Size | PDF

Plane waves are acoustic waves that deliver most of their energy in a forward and backward direction. The initial pressure rise due to these waves can be described by equation 4. In order to obtain plane waves in photoacoustics the following equation must be satisfied: z=d2μa8 where d is the laser beam diameter and z is the acoustic near field/far field boundary as discussed by Viator [13]. In the PAS regime this is generally followed since the pigment of the dyes combined with the low penetration depth along with the geometry of the incoming beam create these conditions. However, with myoglobin, the solution is much more transparent and therefore, plane waves are not necessarily always obtained at every angle of incidence. In the TIRPAS regime, plane waves are always obtained due to the penetration depth of the optical radiation being so small compared to the relative size of the beam, 500 nm vs. 1 mm beam size or more. This can create issues when peak to peak voltage is compared in the PAS and TIRPAS regimes since it is possible for the TIRPAS regime to have similarly sized peak to peak voltage as the PAS regime, despite the PAS regime having more optical penetration depth due to the differences between comparing plane waves in the TIRPAS regime to non-plane waves in the PAS regime. Our data with myoglobin shown in Fig. 6 suggest that this had an effect on our refractive index resolution since the acoustic waves from the PAS region with myoglobin do not show the typical plane wave configuration due to the low concentration of myoglobin used. An alternative method based upon integrated pressure may be more appropriate in the future.

Beyond the peak to peak differences, the two signals’ relative shapes are worth mentioning. The myoglobin wave’s complex shape comes from the fact that a cylindrical segment of sample is being excited that is giving rise to the acoustic wave. First, since angles of incident light smaller than the critical angle are used for the PAS regime, there is an effectively planar wave generated for both Direct Red and myoglobin when the refracted light transfers into the sample region at an angle close to 90 degrees from normal to the interface. This effect is due to the fact that the optical penetration depth is limited since the light is traveling nearly perpendicular to the sample/prism interface. However, at angles of incidence further away from the critical angle in the PAS regime, as shown in Fig. 6, the myoglobin’s signal starts to resemble a wave with more acoustic diffraction due to larger boundary waves being generated from the laser pulse incident at a lower angle of incidence from normal to the interface.

5. Conclusion

As can be seen from table 2, PAS/TIRPAS refractometry should have several uses for materials that are heavily opaque with unknown optical properties. This suggests the use of PAS/TIRPAS refractometry with metals as well as less homogenous materials for quick and simple critical angle analysis. PAS/TIRPAS refractometry has two major advantages over current methods that include the ability to analyze deeply opaque materials along with a true single wavelength determination of the refractive index without the need for complex model fitting needed for other techniques.

Tables Icon

Table 2. Comparison of current technologies

PAS/TIRPAS refractometry has shown promise in its ability to determine refractive index values for myoglobin at a diagnostically relevant concentration. This opens the technique to biological monitoring of myoglobin levels before and after surgery to aid doctors in the diagnosis of potential renal problems occurring from the accumulation of myoglobin after acute hyper-thermia or myocardial infarction. In addition, PAS/TIRPAS refractometry, if combined with an automated syringe pump, could be used for the monitoring of levels of important biological analytes in blood, urine, and for the determination of binding kinetics similar to surface plasmon resonance.

We have demonstrated a new photoacoustic technique that can be used in place of ellipsometry and optical reflectometry for the determination of the real part of the refractive index of liquid solutions. Additionally, we have demonstrated the refractometer’s capability to determine the refractive index of myoglobin at a diagnostically relevant concentration. This new technology has been shown to complement standard technologies such as ellipsometry and reflectometry for future biomedical testing.

6. Appendix 6.1. Stineman smoothing algorithm from Kaleidagraph 4.1.0

The smoothing function in Kaleidagraph’s software is based upon a combination of the Stineman interpolation method [48] with a weighting function. The step by step method is given below:

  1. Fit the Stineman interpolation function to the data.
  2. Apply a weighting function to the data to smooth the curve.
The weighting function takes into account the surrounding 10% of the data points during smoothing which can be understood better by a plot of the distribution in Fig. 7. This is only true if 10% or more of the data is remaining on either side. If less is remaining than 10% of the data the function uses all data that is left until at the final point on the edge of the dataset the value of the Stineman interpolation is used.

 figure: Fig. 7

Fig. 7 Weighting structure for smooth function

Download Full Size | PDF

An array is then created using the following code:

  • coeff[wsize-1]=1.0
  • coeff[kk]=coeff[kk+1]*magic
where wsize is 10% of the number of data points, magic=110abs(wsize) and kk goes from 0 to (wsize-2). From there the weighting function is applied to the data and the smooth curve is output to the program.

Acknowledgments

The authors would like to acknowledge Bobbi J. Hauptmann and Ronald G. Phillips of the Boone County Lumber Building Technology Laboratory within the Department of Architectural Studies at the University of Missouri-Columbia for technical assistance. The authors would like to acknowledge Gordon Ellison for machining the components and for his design expertise. The authors would like to acknowledge Professor Teruo Hinoue for his email correspondence about the history of TIRPAS development. This research was funded in part by a National Science Foundation BRIGE Award ( 1221019). We thank Christine Brethorst for her assistance in performing experiments and analysis of this work

References and links

1. O. Sydoruk, O. Zhernovaya, V. Tuchin, and A. Douplik, “Refractive index of solutions of human hemoglobin from the near-infrared to the ultraviolet range: Kramers-kronig analysis,” J. Biomed. Opt.17(2012). [CrossRef]   [PubMed]  

2. A. S. Laurence, G. K. Vanner, W. Collins, and P. M. Hopkins, “Serum and urinary myoglobin following an aborted malignant hyperthermia reaction,” Anaesthesia 51, 958–961 (1996). [CrossRef]   [PubMed]  

3. A. S. Laurence, “Serum myoglobin and creatine kinase following surgery,” Brit. J. Anaesth. 84, 763–766 (2000). [CrossRef]   [PubMed]  

4. E. Hecht, Optics (4th Edition) (Addison Wesley, 2001), 4th ed.

5. N. J. Harrick, Internal Reflection Spectroscopy (Harrick Scientific Corporation, 1987).

6. R. Paschotta, “Rp photonics encyclopedia,” (2013).

7. A. Rosencwaig and A. Gersho, “Theory of the photoacoustic effect with solids,” J. Appl. Phys. 47, 64–69 (1976). [CrossRef]  

8. A. Rosencwaig, “Photoacoustic spectroscopy,” Annu. Rev. Biophys. Bio. 9, 31–54 (1980). [CrossRef]  

9. L. Wang, “Multiscale photoacoustic microscopy and computed tomography,” Nat. Photonics 3, 503–509 (2009). [CrossRef]  

10. D. McCormack, K. Bhattacharyya, R. Kannan, K. Katti, and J. Viator, “Enhanced photoacoustic detection of melanoma cells using gold nanoparticles,” Laser. Surg. Med. 43, 333–338 (2011). [CrossRef]  

11. C. Hernandez, T. Murray, and S. Krishnaswamy, “Photoacoustic characterization of the mechanical properties of thin films,” Appl. Phys. Lett. 80, 691–693 (2002). [CrossRef]  

12. J. B. Kinney and R. H. Staley, “Applications of photoacoustic spectroscopy,” (Annu. Rev. Inc., Palo Alto, Calif, USA, 1982), vol. 12, pp. 295–321.

13. J. Viator, S. Jacques, and S. A. Prahl, “Depth profiling of absorbing soft materials using photoacoustic methods,” IEEE J. Sel. Top. Quant. 5, 989–996 (1999). [CrossRef]  

14. B. Goldschmidt, A. Sudduth, E. Samson, P. Whiteside, K. Bhattacharyya, and J. Viator, “Total internal reflection photoacoustic spectroscopy for the detection of beta-hematin,” J. Biomed. Opt. 17, 061212 (2012). [CrossRef]  

15. SPIE, Total internal reflection photoacoustic detection spectroscopy, 7899.

16. T. Iwasaki, T. Sawada, H. Kamada, A. Fujishima, and K. Honda, “Observation of semiconductor electrode-dye solution interface by means of fluorescence and laser-induced photoacoustic spectroscopy,” J. Phys. Chem-US 83, 2142–2145 (1979). [CrossRef]  

17. T. Hinoue, Y. Shimahara, and Y. Yokoyama, “Photoacoustic observation of solid-liquid interface by means of total internal reflection technique,” Chem. Lett. 12, 225–228 (1983). [CrossRef]  

18. P. R. Muessig and G. J. Diebold, “Total internal reflectance optoacoustic spectroscopy,” J. Appl. Phys. 54, 4251–4253 (1983). [CrossRef]  

19. T. Hinoue, Y. Shimahara, H. Murata, and Y. Yokoyama, “Photoacoustic spectrometry by total internal reflection technique: dependence of photoacoustic signal intensity on concentration and optical path,” Bunseki Kagaku 33, E459–E466 (1984). [CrossRef]  

20. T. Hinoue, H. Murata, M. Kawabe, and Y. Yokoyama, “Effects of thermal diffusion and solvent materials on photoacoustics signals in total internal reflection technique,” Anal. Sci. 2, 407–410 (1986). [CrossRef]  

21. T. Inagaki, J. Goudonnet, P. Royer, and E. Arakawa, “Optical properties of silver island films in the attenuated-total-reflection geometry,” Appl. Optics 25, 3635–3639 (1986). [CrossRef]  

22. T. Hinoue, H. Murata, and Y. Yokoyama, “Photoacoustic spectrometry coupled with total internal reflection technique: theory and experiment,” Anal. Sci. 2, 401–406 (1986). [CrossRef]  

23. T. Hinoue, M. Kawabe, and Y. Yokoyama, “Measurement of concentration profile of dye at glass-solution interface by photoacoustic spectrometry coupled with total internal reflection technique,” Chem. Soc. Japan 60, 3811–3813 (1987). [CrossRef]  

24. J. Goudonnet, T. Inagaki, E. Arakawa, and T. Ferrell, “Angular and polarization dependence of surface-enhanced raman scattering in attenuated-total-reflection geometry,” Phys. Rev. B 36, 917–921 (1987). [CrossRef]  

25. P. Royer, J. P. Goudonnet, T. Inagaki, G. Chabrier, and E. T. Arakawa, “Photoacoustic study of the optical absorption of oblate silver spheroids in attenuated-total-reflection geometry,” Phys. Status Solidi A 105, 617–625 (1988). [CrossRef]  

26. T. Hinoue, M. Kawabe, S. Doi, and Y. Yokoyama, “Photoacoustic estimation of reflectivities at solid-liquid interfaces by using total internal reflection technique,” J. Surf. Sci. Soc. Japan 10, 47–52 (1989).

27. T. Inagaki, M. Motosuga, E. Arakawa, and J. Goudonnet, “Coupled surface plasmons in periodically corrugated thin silver films,” Phys. Rev. B 32, 6238–6245 (1985). [CrossRef]  

28. S. Negm and H. Talaat, “Effect of intrinsic surface roughness and other decay processes on surface plasmon polariton resonance halfwidth,” Ultrason. 1, 509–514 (1992).

29. B. Rothenhausler, J. Rabe, P. Korpiun, and W. Knoll, “On the decay of plasmon surface polaritons at smooth and rough ag-air interfaces: A reflectance and photo-acoustic study,” Surf. Sci. 137, 373–383 (1984). [CrossRef]  

30. K. Sathiyamoorthy, J. Joseph, C. J. Hon, and M. V. Matham, “Photoacoustic based surface plasmon resonance spectroscopy: an investigation,” pp. 80010K–80010K–8 (2011).

31. K. Sathiyamoorthy, J. Joseph, C. Hon, and M. Matham, “Photoacoustic based surface plasmon resonance spectroscopy: An investigation,” (Braga, 2011), vol. 8001.

32. H. Talaat, J. Bucaro, W. Huang, and A. Macdiarmid, “Photoacoustic detection of plasmon surface polaritons in heavily doped polyacetylene films,” Synthetic Met. 10, 245–253 (1985). [CrossRef]  

33. C. S. Jung, G. Park, and Y. D. Kim, “Photoacoustic determination of field enhancement at a silver surface arising from resonant surface plasmon excitation,” Appl. Phys. Lett. 47, 1165–1167 (1985). [CrossRef]  

34. T. Inagaki, Y. Nakagawa, E. Arakawa, and D. Aas, “Photoacoustic determination of radiative quantum efficiency of surface plasmons in silver films,” Phys. Rev. B 26, 6421–6430 (1982). [CrossRef]  

35. T. Inagaki, K. Kagami, and E. Arakawa, “Photoacoustic observation of nonradiative decay of surface plasmons in silver,” Phys. Rev. B 24, 3644–3646 (1981). [CrossRef]  

36. Y. Jiang, S. Zhang, F. Qian, and H. Shao, “Photoacoustic plasmon resonance spectra of langmuir-blodgett films,” (Editions de Physique, Les Ulis, France, Guadeloupe, Fr, 1994), vol. 4, pp. C7–35–38.

37. Y. S. Jiang, S. Y. Zhang, F. Qian, and H. P. Shao, “Photoacoustic plasmon resonance spectra of langmuir-blodgett films,” J. Phys. IV France 04, C735–C738 (1994). [CrossRef]  

38. T. Inagaki, K. Kagami, and E. Arakawa, “Photoacoustic study of surface plasmons in metals,” Appl. Optics 21, 949–954 (1982). [CrossRef]  

39. S. Negm, H. Talaat, and J. Pelzl, “Photoacoustic study of surface roughness on thin evaporated gold films,” (Publ by IEEE, Piscataway, NJ, United States, Baltimore, MD, USA, 1993), vol. 2, pp. 1259–1261.

40. H. Talaat and H. Dardy, “Photoacoustic study of the interaction of surface plasmons and monolayers of dye molecules,” (IEEE, New York, NY, USA, Atlanta, Ga, USA, 1983), vol. 2, pp. 700–703.

41. T. Abdallah, S. Negm, and H. Talaat, “Photoacoustic surface plasmon for the detection of nicotine,” Egypt. J. Sol. 25, 181–189 (2002).

42. S. Negm and H. Talaat, “Radiative and non-radiative decay of surface plasmons in thin metal films,” Solid State Commun. 84, 133–137 (1992). Cited By (since 1996) 5. [CrossRef]  

43. S. Negm and H. Talaat, “Surface plasmon resonance halfwidths as measured using attenuated total reflection, forward scattering and photoacoustics,” J. Phys.-Condens. Mat. 1, 10201–10205 (1989). [CrossRef]  

44. Y. Shen, S. Zhang, Y. Jiang, R. Zhu, and Y. Wei, “Angular resonance absorption spectra of langmuir-blodgett films studied by the photoacoustic technique,” Thin Solid Films 248, 36–40 (1994). [CrossRef]  

45. M. Xu, S. Zhang, and T. Inagaki, “Investigation of optical resonance absorption on bigratings by photoacoustic angular spectroscopy,” Shengxue Xuebao/Acta Acust.. 25, 440–444 (2000). Cited By (since 1996) 0.

46. Y. Jiang, S. Zhang, H. Shao, and C. Yuan, “Optical properties of langmuir-blodgett films investigated by a photoacoustic technique,” Appl. Optics 34, 169–173 (1995). [CrossRef]  

47. Ultrasonics International, Photoacoustic resonance absorption spectra of langmuir-blodgett films.

48. R. W. Stineman, “A consistently well-behaved method of interpolation,” Creative Comput. pp. 54–57 (1980).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 Schematic of setup showing how laser light interrogates sample to give acoustic effect in TIRPAS
Fig. 2
Fig. 2 Figure illustrates the vastly different optical absorption depths between the PAS and TIRPAS excitation conditions.
Fig. 3
Fig. 3 3-dimensional model of the setup
Fig. 4
Fig. 4 Refractometer setup used to compare refractive index with PAS/TIRPAS refractometry
Fig. 5
Fig. 5 Graphs a, b, c, and d show the step by step data analysis process to determine the critical angle.
Fig. 6
Fig. 6 Acoustic differences between dye solution and myoglobin solution in the PAS regime
Fig. 7
Fig. 7 Weighting structure for smooth function

Tables (2)

Tables Icon

Table 1 PAS/TIRPAS refractometry as compared to a handheld refractometer demonstrates that our method is in good agreement with standard techniques.

Tables Icon

Table 2 Comparison of current technologies

Equations (11)

Equations on this page are rendered with MathJax. Learn more.

n R = sin θ i [ 1 + ( 1 ρ 1 + ρ ) 2 tan 2 θ i ] 1 2
ε ˜ = sin 2 ϕ [ 1 + ( 1 ρ 1 + ρ ) 2 tan 2 ϕ ]
ε ˜ = n 2 = κ
p 0 ( z ) = μ a Γ H 0 e μ a z
μ a = 1 δ
n 1 sin θ 1 = n 2 sin θ 2
θ c = sin ( n 2 n 1 ) 1
E | | = ( E x 2 + E z 2 ) 1 2
E x = 2 E | | ( sin 2 θ n 21 2 ) 1 2 cos θ ( 1 n 21 2 ) 1 2 [ ( 1 + n 21 ) sin 2 θ n 21 2 ] 1 2
E z = 2 E | | sin θ cos θ ( 1 n 21 2 ) 1 2 [ ( 1 + n 21 ) sin 2 θ n 21 2 ] 1 2
δ p = λ 1 2 π ( sin 2 θ n 21 2 ) 1 2
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.