Open Access
8 November 2013 Study of dynamic pressure-induced refractive index change using derivative total reflection method
Jin Wang, Qing Ye, Zhichao Deng, Teng-Qian Sun, Wen-Yuan Zhou, Jian-Chun Mei, Chun-Ping Zhang, Jianguo Tian
Author Affiliations +
Abstract
We report the dynamic refractive index (RI) change of tissue under a stepped compression load using a custom-built pressure apparatus. Angle-dependent reflectance profiles of biotissue samples are recorded, and the RI values are resolved using the derivative total reflection method. These results are relevant for understanding the mechanism of mechanical optical clearing, for investigating tissue dynamics under mechanical stimuli, and for other biomedical applications.

1.

Introduction

The concept of reducing the optical scattering in tissue to enhance the effectiveness of diagnostic and therapeutic treatments has stimulated research in the optical clearing (OC) technique since 1997.1,2 Previous studies have mainly focused on the use of biocompatible and osmotically active chemical agents.3,4 As an important candidate mechanism of OC technique, some researchers shed light on another alternative nonchemical technique of mechanical compression in recent years, which creates fast local effects, maintains stratum corneum integrity and may be harnessed as a minimally invasive OC technique.58 In the previous work, laser beam profiles in response to mechanical compression were measured with experiments carried out on uniform compression between two glass slides, so the compressive force applied to the tissue was unknown.5 Other techniques have been applied for evaluating the effect of OC under direct mechanical compression, such as image analysis,6 optical coherence tomography,7 and diffuse reflectance spectroscopy.8 However, these previous studies have largely focused on transmission or reflection light intensity measurements. A direct and systematic evaluation of mechanical OC effect has not yet been achieved. As the most direct parameter for OC effect evaluation, the change in refractive index (RI) as a function of pressure is required.

On the other hand, for optical properties measurements of biotissue, an appropriate means for applying and maintaining pressure during experiments is required to keep surface tissues immobile, to achieve precise and reproducible treatment of the tissue, and to minimize any related noise contributions.9 In one study involving several laboratories,1012 reflectance measurements and the inverse double method were used to find the relationship between the optical parameters (mainly the reduced scattering and absorption coefficients) and the pressure, whereas a direct study of RI was not attempted. The RI of tissue, as an indicator of the presence of tumors, has gained significant attention in recent years.13 The study of tissue dynamics under mechanical stimuli has provided another impetus to use pressure-induced tumor structural and functional dynamic characteristics for the detection of cancer.14 Knowledge of the change in the dynamic RI of biotissue over a wide pressure range is crucial for these reasons.

The pioneering work of Ding et al. first reported a change in RI with respect to skin sample pressure for applied pressures of <0.55MPa.15 In this article, we report measurements of the change in the dynamic RI of tissue under a stepped compression load (0 to 2.1 MPa) using a custom-built pressure apparatus. Angle-dependent reflectance intensity profiles were measured, and dynamic RI values of tissue were resolved under different pressures using the derivative total reflection method (DTRM). The relationship between the change in RI, the reduction of tissue fluid, and applied pressure is reported. These results are valuable for the evaluation of the mechanical OC effect and provide previously unreported RI values for tissue under applied pressure.

2.

Materials and Methods

The experimental setup used in this work is shown in Fig. 1 and is based on our previously reported work.16 The setup consists of a He-Ne laser (632.8 nm), rotation stage (M-038, Physik Instrumente, Germany), half-wave plate H, polarizer P, aperture diaphragm D, and dual-channel power meter (PM320E, Thorlabs, New Jersey, USA) with two detectors (PD1 and PD2). The effect of fluctuations in the laser power is minimized by the beam splitter M and by the continuous monitoring of the laser power by PD1. Our custom-built pressure apparatus consists of an air pump, a piston-cylinder, and an air gauge to monitor the pressure. An aluminum plate attached to a reciprocating piston can move forward and backward to generate stepped compression loads of 0.35, 0.70, 1.05, 1.4, 1.75, and 2.1 MPa. Samples filled in a plastic groove were sandwiched between the aluminum plate and an equilateral prism and then fixed on the rotation stage. With the aid of a Mercury C-863 servo motor controller and the PI General Command Set software on a computer, the prism was rotated with different incident angles. The reflectance intensity profile collected by PD2 was recorded as a function of incident angle. The illuminated area is inside a circle of radius about 3 mm. Porcine and bovine leg muscles were obtained from the local abattoir immediately after slaughter. Fresh porcine muscle tissue was first frozen for 30 min and then trimmed to a thin section of about 1.2 cm radius and 2 mm thickness to fill the plastic groove. Each sample was measured for ambient and increasing pressures. Bovine tissue was prepared in the same way. The reflectance profile at each pressure was measured three times, and the calculated variation was found to be <3% for all measurements.

Fig. 1

Schematic diagram of the experimental setup.

JBO_18_11_117005_f001.png

Previously, we found that there is a random distribution of tissue fluid and muscle tissue at the prism-tissue interface, while the RI information of tissue fluid and the muscle tissue can be extracted by reflectance measurement, respectively.16 Either the tissue fluid or the muscle tissue can be considered to be homogeneous and to yield an average RI value. The measured reflectance is a result of two types of contact areas. Taking into account the tissue fluid, the reflectance is expressed as

Eq. (1)

Rs+=bfRf+bmRm,
where Rf and Rm are the reflectance from tissue fluid and muscle tissue surface, respectively; bf and bm are fractions of the tissue fluid and the muscle tissue, respectively, at the illuminated areas, and bf+bm=1.

According to the Fresnel formula,17 for transverse electric (TE) wave, Rf is given by

Eq. (2)

Rf=(1R1,2)2×R2,3,

Eq. (3)

R1,2={cosαn1cos[asin(sinα/n1)]cosα+n1cos[asin(sinα/n1)]}2,

Eq. (4)

R2,3={n1cosθnfcos[arcsin(n1sinθ/nf)]n1cosθnfcos[arcsin(n1sinθ/nf)]}2,
where R1,2 and R2,3 denote the reflectance at the air-prism and prism–sample interface, respectively. nf is the RI of tissue fluid. Rm can be calculated in a similar way by replacing nf with the RI of muscle tissue nm in Eqs. (2)–(4). According to the schematic diagram in Fig. 1, nf and nm can be solved by nf=n1sin[β±asin(sinαc,f/n1)] and nm=n1sin[β±asin(sinαc,m/n1)], respectively. As described above, a binary distribution of tissue fluid and muscle should result in a double peak on the derivative curve. αc,f and αc,m are the critical incident angles when total internal reflection occurs at the prism-liquid and prism-tissue interface, respectively, which correspond to the angular positions of the first and second maxima of the derivative curve. Similar equations exist for transverse magnetic (TM) wave. The amplitude of the first and second peak on the reflectance curve indicates the proportion of tissue fluid and muscle tissue in the illuminated area, respectively.16

The model described by Eq. (1) was fitted to a reflectance profile where b1 and b2 are fitting parameters. A nonlinear least-squares fitting program based on the Nelder–Mead simplex algorithm was implemented by MATLAB software, which minimizes the sum S(N)=i=1N[Rm,iRs,i+]2. Rm,i is the i’th measured reflectance. Rs,i+ is the i’th calculated reflectance obtained by Eq. (1). The reliability of the fit is given by Es2, defined as Es2=1i=1N(Rm,iRs,i+)2/i=1N(Rm,iR¯)2, where R¯ is the mean value of measured reflectance over N values of incident angle. Es2 ranges from 0 to 1, where a value near 1 represents a reliable fit.

3.

Results and Discussion

Figure 2(a) shows the measured reflectance data and the model fit for porcine tissue at different pressures. When pressure was applied to the sample, the RI of tissue fluid could be calculated using the value for the critical angle αc,f, which is about 51.8 deg. The amplitude of the first peak on the reflectance curve indicates the proportion of tissue fluid in the illuminated area. The derivative of the reflectance curves is shown in Fig. 2(b). The amplitudes of the first peak gradually disappear as the pressure increases. On account of the pressure effect, the loss of tissue fluid at the prism–tissue interface results in a decrease in the amplitude of the first peak. A large increase of αc,m was observed, and nm significantly increased from 1.373 to 1.450. The RIs of muscle and tissue fluid as a function of the pressure are shown in Fig. 2(c). nf ranges from 1.335 to 1.352 under compression. No noticeable change in nf was observed with an increase in pressure, whereas a noticeable increase in nm was observed when the pressure was higher than 0.35 MPa. The proportions of tissue fluid are calculated and shown in Fig. 2(d). The proportion of tissue fluid reduced from 57.9 to 16.9% when the applied pressure was increased from 0 to 2.1 MPa. The calculated nr of 1.373 at 0 MPa is similar to the reported values of porcine tissue.18

Fig. 2

(a) Measured reflectance data and fitting curves for porcine tissue under different pressure. (b) The derivative of the reflectance curves. (c) Refractive indices (RIs) of tissue fluid and muscle tissue versus the applied pressure. (d) Proportions of tissue fluid under different pressure.

JBO_18_11_117005_f002.png

The RIs of bovine muscle and tissue fluid as a function of the pressure are shown in Fig. 3(a). nf ranged from 1.335 to 1.357 under compression, and nm increased from 1.382 to 1.434. The measured result under the condition of zero pressure is in close agreement with values reported.19 In Fig. 3(b), an obvious reduction in tissue fluid can be observed at pressures >0.35MPa, which resembles the porcine tissue results. The proportion of tissue fluid bf decreased from 79.5 to 25.9% for an increase in applied pressure from 0 to 0.35 MPa. When the pressure increased to 2.1 MPa, only about 6.9% of the total tissue fluid remained.

Fig. 3

(a) RIs of tissue fluid and bovine tissue versus the applied pressure. (b) Proportions of tissue fluid under different pressure.

JBO_18_11_117005_f003.png

We define a relative displacement of tissue fluid Δm to record the loss of tissue fluid under different pressures. Δm=[m(t)m(0)]/m(0), m(0) is the initial weight of tissue sample, m(t) is the final weight of sample after a certain pressure was applied for 5 min. Here, 5 min is the average time in the reflectance measurement for each pressure. The result of Δm is shown in Fig. 4. We can find that the displacement of bovine tissue is much faster than the porcine tissue when the pressure first reached 0.35 MPa. Both the weight of the porcine and bovine tissues reduced to 50% when the pressure reached 2.1 MPa.

Fig. 4

Relative displacement of tissue fluid with different pressure.

JBO_18_11_117005_f004.png

By comparing the results for porcine and bovine tissue, we can specify two critical pressures (P1 and P2) that illustrate the change in RI under pressure. With the condition of no pressure, the tissue fluid on the surface of the sample has a higher RI because of evaporation and dehydration. When the applied pressure reached the first critical pressure P1 (in our experiment, 0.35MPa), the tissue fluid on the surface of the prism flowed away parallel to the surface of the sample, so a clear decrease of the proportion of tissue fluid was observed. The inner fluid of the sample flowed out and diluted the surface fluid, and the RI of tissue fluid decreased, too. The RI of tissue continually increased with the applied pressure. When the applied pressure reached the second critical pressure P2 (in our experiment, 1.4MPa), an increase in the proportion of tissue fluid was observed, which may be attributed to the rupture of cell membranes and the exudation of intracellular fluid. After this increase, both nf and nm increased, while the amount of tissue fluid reduced to a minimal level. A slight change in nf and a continuous increase in nm were observed during the complete compression process. The initial proportion of tissue fluid had a close relationship to sample preparation. The condensation of water on the sample surface during the sample defrosting process was inevitable, and hence, the proportion of tissue fluid at 0 MPa was not a good indication of the intrinsic sample condition. From 0.35 to 2.1 MPa, the speed of the reduction in tissue fluid at a constant pressure for bovine tissue is considerably faster than that for porcine tissue. This observation implies that water loss for bovine tissue under pressure occurs much more readily.

The disparity in the RI values under different compressions indicates that the tissue sample preparation, condition of the prism-sample interface, and intrinsic character of the tissue sample (water content, texture, hardness, and so on) have a significant impact on the final result. The deformation of the cell, displacement of water from collagen fibril, the increase of protein, sugar, and local chromophore concentrations, all contribute to the increase in the RI of muscle tissue under compression.1012

This result also provides an evidence for the RI-matching mechanism of mechanical OC technique from another viewpoint. Since scattering arises from a mismatch in RI, organelles and subcomponents of organelles that have a different RI from their surroundings are expected to be the primary sources of cellular scattering.20 Under certain amounts of compression, an obvious reduction in tissue fluid at the prism-tissue interface yields a higher RI in the contact area, which leads to mechanical OC. This study has also verified that dehydration is an important candidate mechanism for OC and can lead to the matching of the RI for intrinsic structures.7

4.

Conclusion

In conclusion, the amount of pressure applied to the tissue sample affects the RI of muscle tissue and the amount of tissue fluid. An analytical model based on DTRM enables the simultaneous extraction of the change in RI and the proportion of tissue fluid at the contact area. The RI value can be regarded as one important parameter for selecting the proper experimental condition and for evaluating the effect of mechanical OC. In this work, we also report the quantitative values of the RI of compressed biotissue. Considering the recent interest in studies on tissue dynamics under mechanical stimuli, we expect to be able to expand this work by studying more tissue samples with respect to tissue dynamics, optical diagnosis, and treatment.

Acknowledgments

This research is supported by the Chinese National Key Basic Research Special Fund (Grant No. 2011CB922003), Science and Technology Program of Tianjin (Grant No. 11ZCKFSY00500).

References

1. 

V. V. Tuchinet al., “Light propagation in tissues with controlled optical properties,” J. Biomed. Opt., 2 (4), 401 –417 (1997). http://dx.doi.org/10.1117/12.281502 JBOPFO 1083-3668 Google Scholar

2. 

E. A. GeninaA. N. BashkatovV. V. Tuchin, “Tissue optical immersion clearing,” Expert Rev. Med. Devices., 7 (6), 825 –842 (2010). http://dx.doi.org/10.1586/erd.10.50 1743-4440 Google Scholar

3. 

G. Vargaset al., “Use of an agent to reduce scattering in skin,” Lasers Surg. Med., 24 (2), 133 –141 (1999). http://dx.doi.org/10.1002/(ISSN)1096-9101 LSMEDI 0196-8092 Google Scholar

4. 

R. K. Wanget al., “Concurrent enhancement of imaging depth and contrast for optical coherence tomography by hyperosmotic agents,” J. Opt. Soc. Am. B, 18 (7), 948 –953 (2001). http://dx.doi.org/10.1364/JOSAB.18.000948 JOBPDE 0740-3224 Google Scholar

5. 

H. Kanget al., “Evaluation of laser beam profile in soft tissue due to compression, glycerol, and micro-needling,” Lasers Surg. Med., 40 (8), 570 –575 (2008). http://dx.doi.org/10.1002/lsm.v40:8 LSMEDI 0196-8092 Google Scholar

6. 

A. Izquierdo-Románet al., “Erratum: mechanical tissue optical clearing technique increases imaging resolution and contrast through ex vivo porcine skin,” Lasers Surg. Med., 43 (10), 814 –823 (2011). http://dx.doi.org/10.1002/lsm.21105 LSMEDI 0196-8092 Google Scholar

7. 

C. DrewT. E. MilnerC. G. Rylander, “Mechanical tissue optical clearing devices: evaluation of enhanced light penetration in skin using optical coherence tomography,” J. Biomed. Opt., 14 (6), 064019 (2009). http://dx.doi.org/10.1117/1.3268441 JBOPFO 1083-3668 Google Scholar

8. 

W. C. Vogtet al., “Effects of mechanical indentation on diffuse reflectance spectra, light transmission, and intrinsic optical properties in ex vivo porcine skin,” Lasers Surg. Med., 44 (4), 303 –309 (2012). http://dx.doi.org/10.1002/lsm.v44.4 LSMEDI 0196-8092 Google Scholar

9. 

J. Chaikenet al., “Instrument for near infrared emission spectroscopic probing of human fingertips in vivo,” Rev. Sci. Instrum., 81 (3), 034301 (2010). http://dx.doi.org/10.1063/1.3314290 RSINAK 0034-6748 Google Scholar

10. 

H. Shangguanet al., “Pressure effects on soft tissues monitored by changes in tissue optical properties,” Proc. SPIE, 3254 366 –371 (1998). http://dx.doi.org/10.1117/12.308187 PSISDG 0277-786X Google Scholar

11. 

E. K. Chanet al., “Effects of compression on soft tissue optical properties,” IEEE J. Sel. Top. Quantum Electon., 2 (4), 943 –950 (1996). http://dx.doi.org/10.1109/2944.577320 IJSQEN 1077-260X Google Scholar

12. 

R. Reifet al., “Analysis of changes in reflectance measurements on biological tissues subjected to different probe pressures,” J. Biomed. Opt., 13 (1), 010502 (2008). http://dx.doi.org/10.1117/1.2870115 JBOPFO 1083-3668 Google Scholar

13. 

Z. Wanget al., “Tissue refractive index as marker of disease,” J. Biomed. Opt., 16 (11), 116017 (2011). http://dx.doi.org/10.1117/1.3656732 JBOPFO 1083-3668 Google Scholar

14. 

B. Wanget al., “Dynamic schema for near infrared detection of pressure-induced changes in solid tumors,” Appl. Opt., 47 (16), 3053 –3063 (2008). http://dx.doi.org/10.1364/AO.47.003053 APOPAI 0003-6935 Google Scholar

15. 

H. F. Dinget al., “Refractive indices of human skin tissues at eight wavelengths and estimated dispersion relations between 300 and 1600 nm,” Phys. Med. Biol., 51 (6), 1479 –1489 (2006). http://dx.doi.org/10.1088/0031-9155/51/6/008 PHMBA7 0031-9155 Google Scholar

16. 

J. Wanget al., “Effect of tissue fluid on accurate determination of the complex refractive index of animal tissue,” J. Biomed. Opt., 17 (7), 075011 (2012). http://dx.doi.org/10.1117/1.JBO.17.7.075011 JBOPFO 1083-3668 Google Scholar

17. 

M. BornE. Wolf, Principles of Optics, Pergamon, New York (1959). Google Scholar

18. 

H. LiS. S. Xie, “Measurement method of the refractive index of biotissue by total internal reflection,” Appl. Opt., 35 (10), 1793 –1795 (1996). http://dx.doi.org/10.1364/AO.35.001793 APOPAI 0003-6935 Google Scholar

19. 

J. J. J. DirckxL. C. KuypersW. F. Decraemer, “Refractive index of tissue measured with confocal microscopy,” J. Biomed. Opt., 10 (4), 044014 (2005). http://dx.doi.org/10.1117/1.1993487 JBOPFO 1083-3668 Google Scholar

20. 

R. DrezekA. DunnR. Richards-Kortum, “Light scattering from cells: finite-difference time-domain simulations and goniometric measurements,” Appl. Opt., 38 (16), 3651 –3661 (1999). http://dx.doi.org/10.1364/AO.38.003651 APOPAI 0003-6935 Google Scholar
© 2013 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2013/$25.00 © 2013 SPIE
Jin Wang, Qing Ye, Zhichao Deng, Teng-Qian Sun, Wen-Yuan Zhou, Jian-Chun Mei, Chun-Ping Zhang, and Jianguo Tian "Study of dynamic pressure-induced refractive index change using derivative total reflection method," Journal of Biomedical Optics 18(11), 117005 (8 November 2013). https://doi.org/10.1117/1.JBO.18.11.117005
Published: 8 November 2013
Lens.org Logo
CITATIONS
Cited by 4 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Tissues

Reflectivity

Refractive index

Fluid dynamics

Tissue optics

Natural surfaces

Interfaces

Back to Top