Open Access
15 March 2018 Measurement of refractive index of hemoglobin in the visible/NIR spectral range
Author Affiliations +
Abstract
This study is focused on the measurements of the refractive index of hemoglobin solutions in the visible/near-infrared (NIR) spectral range at room temperature for characteristic laser wavelengths: 480, 486, 546, 589, 644, 656, 680, 930, 1100, 1300, and 1550 nm. Measurements were performed using the multiwavelength Abbe refractometer. Aqua hemoglobin solutions of different concentrations obtained from human whole blood were investigated. The specific increment of refractive index on hemoglobin concentration and the Sellmeier coefficients were calculated.

1.

Introduction

The refractive index (RI) of biological tissue is a basic material parameter that characterizes how light interacts with tissue.1 In many optical studies, a rough estimate of RI of the tissue under study, based on the fact that the main constituent of tissue is salt water-filled cells or more precisely a mixture of salt water and proteins, is often used.2,3 For many tissues and blood components, the data for RI in a wide spectral range and concentrations are not yet available.4

The study of optical properties of hemoglobin is important for the development of diagnostic and laser treatment techniques, where consideration of blood optical properties is critical. Various optical methods widely used for tissue characterization, such as visible and near-infrared (NIR) spectroscopy, optical coherence tomography, and fluorescence spectroscopy, need exact data for RI of tissue, blood, and their components to quantify properly experimental data.511

At present, the usage of RI as a diagnostic marker is urgent.520 Sometimes, it is a self-sufficient parameter for tissue and blood characterization. Zhernovaya et al. considered the change of RI of hemoglobin and albumin at the interaction with glucose as a possible method for studying the glycation process and determining glycated proteins, which is important for the monitoring of diabetes mellitus.9 The RI of tissue was used as a marker for cancer, reflecting changes of optical properties in the course of pathology development.10,11,1420 An additional motivation for this study is a lot of discrepancies between the RIs reported in the literature by different research groups.

The RI is a complex value consisting of a real part n, which represents the ratio of the speed of light in a vacuum to the speed of light in the material n=c/v, and an imaginary part k, which represents light attenuation2125

Eq. (1)

n˜=n+ik.

Because of tissue heterogeneity, n is always known as the effective or average RI.24,26 According to the classical theory of light dispersion, the components of the complex RI of molecular structures can be written as2125

Eq. (2)

n=1+2πq2N(ω02ω2)m(ω02ω2)2+γ2ω2,

Eq. (3)

k=2πq2Nγωm(ω02ω2)2+γ2ω2,
where q is the molecular charge, N is the number of molecules per unit volume, m is the molecular mass, ω is the probing light frequency, ω0 is the central frequency of molecular absorption band, and γ is the attenuation coefficient.2125

Over the last decades, various techniques to determine RI of biological tissues were developed; they include confocal microscopy,1,6 optical fiber cladding method,27 minimum deviation angle method,28,29 optical coherent tomography with multiple modifications,6,9,3036 total internal reflection method,26,37,38 measurement of the intensity profile of diffuse light refracted into the prism around the critical angle,39 various modifications of nonlinear phase microscopy,4042 and quantitative phase imaging techniques.12,43,44

Because of the strong hemoglobin absorption, direct measurements of the real part of RI using conventional refractometers (for example, an Abbe refractometer) have proven to be difficult, and data are available at a few wavelengths only. In an early study, for example, Barer measured n for solutions of oxygenated hemoglobin at 589 nm only.44 He also discussed RI dependence on the hemoglobin concentration and presented the expression

Eq. (4)

n=nH2O+αC,
where nH2O is the RI of distilled water, C is the concentration of hemoglobin, and α is the specific refraction increment.44

Faber et al. measured the RI of solutions of oxygenated and deoxygenated hemoglobin at 800 nm.45 Friebel and Meinke measured directly the RI of solutions of oxygenated hemoglobin at 633 nm for several concentrations7,8

Eq. (5)

n=nH2O(1+βC).

Zhernovaya et al. also used the formula similar to Eq. (4) to describe the linear dependence of the RI of hemoglobin on the concentration

Eq. (6)

n=n0+αC,
where n0 is the RI of solvent, C is the hemoglobin concentration in dl/g, and α is the specific refraction increment.4,46

Jin et al. measured RI of hemoglobin solution at 633 and 532 nm using a total internal reflection technique.37 Park et al. measured the dispersion of Hb solutions, prepared from Hb protein powder, at 440, 546, 560, 580, 600, 655, and 700 nm using spectroscopic phase microscopy.41 Deng et al. showed that, in the 400 to 750 nm range, hemoglobin solution is characterized by specific forms of dispersion and extinction spectra.47 Yahya and Saghir measured RIs for multiple temperatures and wavelengths using the Abbemat refractometer.48 They found linear dependences of RI on hemoglobin concentration and temperature and nonlinear on the wavelength.

Analysis of the dispersion relation in similar studies showed significant differences for oxyhemoglobin and deoxyhemoglobin, related to the difference in the imaginary part of the RI for the 500 to 600 nm region.4,4550 There is lack of data for RI of hemoglobin solutions for concentrations close to that in the red blood cells (RBC), especially for the NIR region.

This study is focused on the determination of the RI of hemoglobin in the visible and NIR ranges at room temperature, aiming for further quantification dispersion of hemoglobin solutions. Measurements were carried out using the multiwavelength Abbe refractometer (Atago, Japan). The hemoglobin solutions of different concentrations obtained from human whole blood were investigated. The RI of hemoglobin solutions was measured for the wavelengths: 480, 486, 546, 589, 644, 656, 680, 930, 1100, 1300, and 1550 nm, which are characteristic for different lasers widely used in biomedicine. The specific increment of RI and Sellmeier coefficients for dispersion on hemoglobin concentration were calculated based on the experimental data.

2.

Methods and Materials

Hemoglobin obtained from human whole blood was used to prepare hemoglobin specimens. Whole blood was drawn from the human vein. Immediately after collecting blood into a test tube, heparin was added in it. The sample of blood from a healthy person was taken at the State Healthcare Organization “Saratov City Clinical Hospital No. 2 named after V. I. Razumovsky” with the permission of the volunteer. To separate blood into fractions, the centrifugation for 10 min at 2000 rpm and at room temperature was provided. This resulted in separation of blood plasma, leuko-platelet layer, and RBC suspension. To conduct hemolysis and preparation of hemoglobin solutions, RBC suspension was separated and placed in a vial for freezing in a freezer at a temperature of 15°C for 24 h.

Actual concentration of the basic hemoglobin solution was estimated by the spectral technique and amounted to 260  g/l. In the experiment, we measured the RI of three specimens taken from the same sample. The RI measurements for solutions of different concentrations obtained by diluting the basic solution of hemoglobin in saline solution were also provided.

At measurements, a sample layer on the working surface of the prism had a small thickness of about 20 to 30  μm. The time of the full oxygenation (78.4% to 94.2%) of hemoglobin in such a layer is about 6 to 10 s.51 Therefore, hemoglobin is fully saturated with oxygen, and the process of oxygenation during measurements is expected to not affect the result.

Measurements were performed using the multiwavelength Abbe refractometer (Atago, Japan) (Fig. 1). The RI was measured for samples of hemoglobin obtained from human whole blood (65, 87, 173, and 260  g/l) on 11 wavelengths from 480 to 1550 nm. The temperature was 23°C.

Fig. 1

General view of the multiwavelength Abbe refractometer (Atago, Japan): 1, refractometer; 2,–power supply; 3, light source; 4, the eyepiece imager for measurements in the NIR region; 5, interference filter; and 6, sample.

JBO_23_3_035004_f001.png

Multiwavelength refractometer Abbe allows one to measure the RI in the wavelength range of 450 to 1550 nm with an accuracy of ±0.0002. The working principle of the refractometer technique is based on determining the critical angle of the total reflection, where the incident light waves are completely reflected with a 90-deg angle to the normal position. The incident light waves with angles greater than the critical angle will only experience reflection at the interface surface and no refraction will be observed. The total internal reflection method is applicable to measurement of the RI of biological media, which are characterized by high light scattering and absorption. The wavelength of the light source is determined by the selection of the particular interferential filter. Available interferential filters allowed for measurements on the wavelengths 480±2, 486±2, 546±2, 589±2, 644±2, and 656±2  nm, 680±5, 930±6, 1100±26, 1300±25, and 1550±25  nm. The calibration of the device by measuring RI of distilled water at a wavelength of 589 nm (the absorption band of sodium) was used at the beginning of each experiment. The average measurement error of the RI was ±0.0003.

To approximate the dispersion dependence of the RI of the hemoglobin solution, the Sellmeier formula was used

Eq. (7)

n2(λ)=1+A1*λ2λ2B1+A2*λ2λ2B2,
where A1, A2, B1, and B2 are empirical constants. Sellmeier’s formula gives a good agreement for describing the dispersion dependence of the RI of multicomponent systems near absorption bands of a medium under study.52 Mathematical calculations were performed in the software package Origin ProLab.

3.

Measurement Results

The optical density spectra of a solution of hemoglobin obtained from the whole blood by hemolysis are shown in Fig. 2. The graph shows that the wavelengths available for RI measurements, i.e., 480, 486, 546, 589, 644, and 656 nm, belong to different or the same absorption bands of hemoglobin with quite different absorption abilities. Therefore, we can expect different inclusion of anomalous dispersion in RI wavelength dependence at these wavelengths. Wavelength 546 nm is the closest to the isobestic point 544 nm, where the absorption of hemoglobin does not depend on the degree of oxygenation.53

Fig. 2

The optical density spectrum of a solution of hemoglobin 260  g/l. By the vertical lines visible working wavelengths of Atago refractometer are shown.

JBO_23_3_035004_f002.png

Table 1 presents data for Atago refractometer measurements of RI for four different concentrations of hemoglobin, i.e., 65, 87, 173, and 260  g/l at room temperature, 23°C.

Table 1

RI measured for four different concentrations of hemoglobin at room temperature 23°C. SD is shown in brackets.

λ (nm)0  g/l65  g/l87  g/l173  g/l260  g/l
4801.3371 (0.0003)1.3476 (0.0003)1.3571 (0.0003)1.3728 (0.0003)1.3879 (0.0002)
4861.3371 (0.0002)1.3478 (0.0002)1.3563 (0.0002)1.3721 (0.0002)1.3871 (0.0004)
5461.3342 (0.0002)1.3448 (0.0002)1.3533 (0.0002)1.3681 (0.0007)1.3836 (0.0002)
5891.3329 (0.0002)1.3438 (0.0002)1.3519 (0.0003)1.3667 (0.0004)1.3821 (0.0004)
6441.3313 (0.0002)1.3419 (0.0002)1.3497 (0.0002)1.3640 (0.0003)1.3801 (0.0003)
6561.3308 (0.0002)1.3414 (0.0002)1.3493 (0.0002)1.3647 (0.0003)1.3792 (0.0009)
6801.3301 (0.0002)1.3403 (0.0003)1.3482 (0.0003)1.3633 (0.0003)1.3771 (0.0002)
9301.3259 (0.0002)1.3360 (0.0002)1.3440 (0.0002)1.3572 (0.0003)1.3735 (0.0007)
11001.3222 (0.0002)1.3329 (0.0002)1.3411 (0.0002)1.3542 (0.0002)1.3690 (0.0006)
13001.3174 (0.0002)1.3280 (0.0005)1.3364 (0.0002)1.3503 (0.0002)1.3642 (0.0004)
15501.3140 (0.0002)1.3244 (0.0004)1.3314 (0.0003)1.3458 (0.0002)1.3598 (0.0004)

It is well known that the RI of proteins is nonlinearly dependent on the wavelength.4,4550,54 Figure 3 shows the dispersion curves for hemoglobin solutions in the visible/NIR spectral range. The symbols are experimental data from Table 1, and the lines correspond to the fit of these data to the Sellmeier formula, Eq. (7). Table 2 presents data for the decomposition of the Sellmeier formula.

Fig. 3

The dispersion curves for hemoglobin solutions: the symbols are experimental data from Table 1 and the lines correspond to the fit of these data to the Sellmeier formula, Eq. (7).

JBO_23_3_035004_f003.png

Table 2

Coefficients of Sellmeier formula for hemoglobin solutions of different concentrations.

Hb (g/l)A1A2B1 (1/nm2)B2, 107 (1/nm2)R2
650.79099685.082378366.452394024.350.995
870.80835450.241199983.697492842.830.999
1730.84507402.8987311065.321172540.720.998
2600.88871190.9531910187.171671039.980.993

As it follows from Table 2, for all wavelengths and hemoglobin concentrations, measured RIs are well fit to the Sellmeier formula with correlation coefficient, R2, equal or better than 0.993. Specifically, there is a linear relationship between the RI and hemoglobin concentration. The RI of the hemoglobin samples is also temperature-dependent, although the temperature effect on the RI is small when compared with the hemoglobin concentration effect. Figure 4 shows the dependence of the RI of human hemoglobin solution on hemoglobin concentration for the room temperature of 23°C. These data can be used to predict the hemoglobin concentration of the blood sample based on the knowledge of the RI and using the refraction increment provided. This dependence can be described by Eqs. (4) and (5).

Fig. 4

The dependence of the RI on the concentration of hemoglobin in solution for: (a) visible and (b) NIR ranges (black symbols, experimental data; red lines, approximation of these data).

JBO_23_3_035004_f004.png

Table 3 presents data for the RI of distilled water and the specific increment of RI for hemoglobin solutions obtained by hemolysis. Approximation of the dependence of the specific increment of the RI on the wavelength was performed using the software package OriginProLab. The best fit was achieved using

Eq. (8)

y=Cx(D+x),
where C=0.17263±0.00157 and D=57.8324±5.56032. The correlation coefficient was R2=0.90.

4.

Discussion

The results of the measurements revealed that there is a linear relationship between the RI and hemoglobin concentration. Table 4 summarizes data on hemoglobin RI available in the literature. The comparison of received data with the literature is presented.

Table 3

The distilled water RI nH2O and the specific increment dn/dC of RI for hemoglobin solutions obtained by hemolysis, for the room temperature 23°C. SD is shown in brackets.

λ (nm)nH2Oα (ml/g)β (ml/g)
4801.3371 (0.0003)0.199 (0.006)0.149 (0.005)
4861.3371 (0.0002)0.196 (0.005)0.147 (0.004)
5461.3342 (0.0001)0.193 (0.005)0.144 (0.004)
5891.3329 (0.0002)0.192 (0.005)0.144 (0.003)
6441.3313 (0.0002)0.189 (0.004)0.142 (0.003)
6561.3308 (0.0002)0.190 (0.005)0.143 (0.003)
6801.3301 (0.0001)0.185 (0.005)0.139 (0.004)
9301.3259 (0.0002)0.183 (0.004)0.138 (0.003)
11001.3222 (0.0002)0.183 (0.005)0.139 (0.004)
13001.3174 (0.0002)0.185 (0.006)0.140 (0.004)
15501.3140 (0.0002)0.179 (0.004)0.136 (0.003)

Table 4

The experimental data for the real part of the RI of the hemoglobin solutions.

λ (nm)g/lNNotesRef.
250461.398Human hemoglobin from fresh RBC suspensions of donors; VIS-NIR-spectrometer55, 56
1041.406
1651.435
2871.470
300461.373
1041.389
1651.405
2871.441
400461.354
1041.367
1651.383
2871.409
400201.35223Bovine hemoglobin (dry); Hb; pH 7.4; room temperature; continuous RI dispersion (CRID)47
401.35495
601.35806
801.36078
1201.36369
1401.36600
2801.37010
3201.38621
400201.35107Bovine hemoglobin (dry); HbO2; pH 7.4; room temperature; CRID
401.35417
601.35767
801.36039
1201.36369
1401.36602
2801.36951
3201.38660
4003201.3822Bovine hemoglobin (lyophilized powder); 0.5% HbO2; T=20°C; fiber spectrometer26
4003201.3775Bovine hemoglobin (lyophilized powder); Hb; T=20°C; fber spectrometer
4011401.365Human hemoglobin (lyophilized powder); Hb; T=20°C; pH 7.4; TIR (total internal reflection)4, 46
435.81.367
4011401.369Human hemoglobin (lyophilized powder); HbO2; T=20°C; pH 7.4;TIR
435.81.366
4361501.36481Human hemoglobin (dry); T=20°C; pH 7.4; Abbemat refractometer48
4381401.374Bovine hemoglobin (dry); Hb; HbO2; room temperature; pH 7.447
440501.3562Human hemoglobin (lyophilized powder); spectroscopic phase microscopy41
1501.3780
3001.4187
4503201.3888Bovine hemoglobin (lyophilized powder); 0.5% HbO2;T=20°C; fiber spectrometer22
4503201.3933Bovine hemoglobin (lyophilized powder); Hb; T=20°C; fiber spectrometer
480651.3476 (0.0003)Human hemoglobin from whole blood; HbO2; T=23°C; Multiwavelength Abbe refractometera
871.3571 (0.0003)
1731.3728 (0.0003)
2601.3879 (0.0002)
486651.3478 (0.0002)Human hemoglobin from whole blood; HbO2; T=23°C; Multiwavelength Abbe refractometera
871.3563 (0.0002)
1731.3721 (0.0002)
2601.3871 (0.0004)
486.11401.361Human hemoglobin (lyophilized powder); Hb; T=20°C; pH 7.4; TIR4, 46
486.11401.361Human hemoglobin (lyophilized powder); HbO2; T=20°C; pH 7.4; TIR
5002871.413Human hemoglobin from fresh RBC suspensions of donors; VIS-NIR-spectrometer55, 56
1651.383
1041.363
461.348
500201.34583Bovine hemoglobin (dry); Hb; pH 7.4; room temperature; CRID47
401.34913
601.35223
801.35592
1201.35922
1401.36175
2801.36544
3201.38408
500201.34505Bovine hemoglobin (dry); HbO2; pH 7.4; room temperature; CRID
401.34854
601.35262
801.35573
1201.35845
1401.36214
2801.36544
3201.38505
513.91501.36053Human hemoglobin (dry); T=20°C; pH 7.4; Abbemat refractometer48
5321.71.3400Human hemoglobin (fresh human blood); T=25°C; TIR37
2.51.3431
41.3485
71.3604
12.971.3871
546651.3448 (0.0002)Human hemoglobin from whole blood; HbO2; T=23°C; Multiwavelength Abbe refractometera
871.3533 (0.0002)
1731.3681 (0.0007)
2601.3836 (0.0002)
546501.3472Human hemoglobin (lyophilized powder); spectroscopic phase microscopy41
1501.3700
3001.4051
546.11401.357Human hemoglobin (lyophilized powder); Hb; T=20°C; pH 7.4; TIR4, 46
546.11401.357Human hemoglobin (lyophilized powder); HbO2; T=20°C; pH 7.4; TIR
5503201.3724Bovine hemoglobin (lyophilized powder); 0.5% HbO2; T=20°C; fiber spectrometer26
5503201.3738Bovine hemoglobin (lyophilized powder); Hb; T=20°C; fiber spectrometer
560501.3466Human hemoglobin (lyophilized powder); spectroscopic phase microscopy41
1501.3687
3001.4033
580501.3451Human hemoglobin (lyophilized powder); spectroscopic phase microscopy41
1501.3668
3001.4025
587.61401.356Human hemoglobin (lyophilized powder); Hb; T=20°C; pH 7.4; TIR4, 46
587.61401.357Human hemoglobin (lyophilized powder); HbO2; T=20°C; pH 7.4; TIR
589651.3438 (0.0002)Human hemoglobin from whole blood; HbO2; T=23°C; multiwavelength Abbe refractometera
871.3519 (0.0003)
1731.3667 (0.0004)
2601.3821 (0.0004)
589461.343Human hemoglobin from fresh RBC suspensions of donors; VIS-NIR-spectrometer55, 56
1041.357
1651.375
2871.406
589.21501.35724Human hemoglobin (dry); T=20°C; pH 7.4; Abbemat refractometer48
589.31401.356Human hemoglobin (lyophilized powder); Hb; T=20°C; pH 7.4;TIR4, 46
589.31401.357Human hemoglobin (lyophilized powder); HbO2; T=20°C; pH 7.4;TIR
600501.3443Human hemoglobin (lyophilized powder); spectroscopic phase microscopy41
1501.3666
3001.4014
600201.34233Bovine hemoglobin (dry); Hb; pH 7.4; room temperature; CRID47
401.34485
601.34874
801.34835
1201.3520
1401.35495
2801.36155
3201.38233
600201.34136Bovine hemoglobin (dry); HbO2; pH 7.4; room temperature; CRID
401.34447
601.34874
801.35068
1201.35456
1401.35767
2801.36155
3201.38058
6003201.3684Bovine hemoglobin (lyophilized powder); 0.5% HbO2; T=20°C; fiber spectrometer26
6003201.3702Bovine hemoglobin (lyophilized powder); Hb; T=20°C; fiber spectrometer
6321.71.3626Human hemoglobin (fresh human blood); T=25°C; TIR37
2.51.3360
41.3425
71.3538
12.971.3800
632.81401.354Human hemoglobin (lyophilized powder); Hb; T=20°C; pH 7.4; TIR4, 46
656.31.354
632.81401.355Human hemoglobin (lyophilized powder); HbO2; T=20°C; pH 7.4; TIR
656.31.354
633.21501.35601Human hemoglobin (dry); T=20°C; pH 7.4; Abbemat refractometer48
657.21.35587
6331041.3600Human hemoglobin (dry); T=20°C; pH 7.4; Abbemat refractometer56
1651.3750
644651.3419 (0.0002)Human hemoglobin from whole blood; HbO2; T=23°C; multiwavelength Abbe refractometera
871.3497 (0.0002)
1731.3640 (0.0003)
2601.3801 (0.0003)
6503201.3652Bovine hemoglobin (lyophilized powder); 0.5% HbO2; T=20°C; fiber spectrometer26
6503201.3668Bovine hemoglobin (lyophilized powder); Hb; T=20°C; fiber spectrometer
655501.3408Human hemoglobin (lyophilized powder); spectroscopic phase microscopy41
1501.3642
3001.3969
656651.3414 (0.0002)Human hemoglobin from whole blood; HbO2; T=23°C; multiwavelength Abbe refractometera
871.3493 (0.0002)
1731.3647 (0.0003)
2601.3792 (0.0009)
680651.3403 (0.0003)Human hemoglobin from whole blood; HbO2; T=23°C; multiwavelength Abbe refractometera
871.3482 (0.0003)
1731.3633 (0.0003)
2601.3771 (0.0002)
700501.3405Human hemoglobin (lyophilized powder); spectroscopic phase microscopy41
1501.3634
3001.3971
700201.33961Bovine hemoglobin (dry); Hb; pH 7.4; room temperature; CRID47
401.34252
601.34602
801.34874
1201.35184
1401.35456
2801.35806
3201.37709
700201.33883Bovine hemoglobin (dry); HbO2; pH 7.4; room temperature; CRID
401.34175
601.34583
801.34835
1201.35107
1401.35476
2801.35748
3201.3767
7003201.3612Bovine hemoglobin (lyophilized powder); 0.5% HbO2; T=20°C; fiber spectrometer26
7003201.3637Bovine hemoglobin (lyophilized powder); Hb; T=20°C; fiber spectrometer
700461.341Human hemoglobin from fresh RBC suspensions of donors; VIS-NIR-spectrometer55, 56
1041.356
1651.374
2871.404
706.51401.352Human hemoglobin (lyophilized powder); Hb; T=20°C; pH 7.4; TIR4, 46
706.51401.352Human hemoglobin (lyophilized powder); HbO2; T=20°C; pH 7.4; TIR
7503201.3589Bovine hemoglobin (lyophilized powder); 0.5% HbO2; T=20°C; fiber spectrometer26
7503201.3599Bovine hemoglobin (lyophilized powder); Hb; T=20°C; Fiber spectrometer
800461.338Human hemoglobin from fresh RBC suspensions of donors; VIS-NIR-spectrometer55, 56
1041.353
1651.370
2871.400
900461.338Human hemoglobin from fresh RBC suspensions of donors; VIS-NIR-spectrometer55, 56
1041.352
1651.369
2871.401
930651.3360 (0.0002)Human hemoglobin from whole blood; HbO2; T=23°C; Multiwavelength Abbe refractometera
871.3440 (0.0002)
1731.3572 (0.0003)
2601.3735 (0.0007)
1000461.338Human hemoglobin from fresh RBC suspensions of donors; VIS-NIR-spectrometer55, 56
1041.353
1651.370
2871.401
1100461.337Human hemoglobin from fresh RBC suspensions of donors; VIS-NIR-spectrometer55, 56
1041.352
1651.369
2871.400
1100651.3329 (0.0002)Human hemoglobin from whole blood; HbO2; T=23°C; Multiwavelength Abbe refractometera
871.3411 (0.0002)
1731.3542 (0.0002)
2601.3690 (0.0006)
1300651.3280 (0.0005)Human hemoglobin from whole blood; HbO2; T=23°C; Multiwavelength Abbe refractometera
871.3364 (0.0002)
1731.3503 (0.0002)
2601.3642 (0.0004)
1550651.3244 (0.0004)Human hemoglobin from whole blood; HbO2; T=23°C; ultiwavelength Abbe refractometera
871.3314 (0.0003)
1731.3458 (0.0002)
2601.3598 (0.0004)

aData from this study.

There is lack of data on the RI measurement of hemoglobin solutions for concentrations close to that in the RBC; specifically, data for the NIR region are practically absent. The RI of hemoglobin solution of 260  g/l, obtained from whole blood at room temperature (23°C) for the wavelength of 480 nm, was found to be equal to 1.3879±0.0002, for 589 nm to 1.3821±0.0004, for 1100 nm to 1.3690±0.0006, and for 1550 nm to 1.3598±0.0002. The concentration increment of RI of hemoglobin was found as 0.199±0.006  ml/g for the wavelength 480 nm, 0.192±0.005  ml/g for the wavelength 589 nm, 0.183±0.005  ml/g for the wavelength 930 nm, and 0.179±0.004  ml/g for the wavelength 1550 nm.

Freibel et al. also measured the RI of a hemoglobin solution of 287  g/l obtained from whole blood. According to their measurements using the spectral method and the Fresnel formula, the RI was 1.409 for the wavelength 400 nm, 1.406 for the wavelength 589 nm, 1.404 for the wavelength 700 nm, and 1.400 for the wavelength 1100 nm.55,56 The same scientific group received at the wavelength 633 nm the RI=1.3750 for concentration 165  g/l and the RI=1.3600 for concentration 104  g/l. Jin et al.,37 Park et al.,41 Zhernovaya et al.,46 Yahya et al.,48 and Deng et al.47 used a solution obtained from dry hemoglobin for the study of refraction. Zhernovaya et al. measured the RI of oxygenated and deoxygenated hemoglobin of 140  g/l by refractometer Abbe for nine wavelengths at a temperature of 20°C. For example, the values of RI were 1.361 for the wavelength 486 nm, 1.357 for the wavelength 589 nm, and 1.352 for the wavelength 706.5 nm.46 Yahya et al. measured RI of oxygenated human hemoglobin 150  g/l as 1.36481 for the wavelength 436 nm, 1.35724 for the wavelength 589 nm, and 1.35587 for the wavelength 657.2 nm.48 Deng et al. measured RI of 50% oxyhemoglobin 320  g/l by fiber spectrometer at a temperature of 20°C. RIs were 1.3775 for the wavelength 500 nm, 1.3684 for the wavelength 600 nm, and 1.3612 for the wavelength 700 nm.47 Jin et al. determined RI of hemoglobin for concentration of 12.97  mmol/l as 1.3871 for the wavelength 532 nm and 1.3800 for the wavelength 632 nm.37 Park et al. measured the dispersion of Hb solutions, prepared from Hb protein powder, at three different concentrations: 0.05, 0.15, and 0.30  g/ml. For example, the RI for 0.15  g/ml was 1.3687 at wavelength 560 nm.

Zhernovaya et al.,46 Freibel et al.,55,56 Yahya et al.,48 and Park et al.41 also calculated the specific increment of RI (20°C), which was equal to 0.147, 0.2015, and 0.151  ml/g for the wavelength 589 nm and 0.183±0.003  ml/g for the wavelength range of 440 to 700 nm, respectively. In this study, the RI-specific increment of hemoglobin was found as 0.192±0.005  ml/g for the wavelength 589 nm and temperature at 23°C.

The discrepancy between literature and our data may be caused by the differences in the sample preparation protocols since the human hemoglobin may differ in content of various forms of hemoglobin of donor’s blood. The specificities of experimental setups also may play a role.

In Fig. 5, it is seen that the RI-specific increment of a solution of human hemoglobin decreases with the wavelength. This could be explained by the dispersion theory of multicomponent materials and caused by strong absorption bands of hemoglobin and water in UV, hemoglobin in the visible, and water in the NIR. The dependence of the specific RI increment α of hemoglobin solution on the wavelength is in a good agreement with the literature data given by Friebel et al. for whole blood using an integrating sphere spectrometer technique and by Jung et al. for an Hb solution in intact individual RBC cytoplasm.43,55

Fig. 5

The dependence of the specific RI increment α of hemoglobin solution on the wavelength (black symbols, estimated data, red lines, approximation of data).

JBO_23_3_035004_f005.png

As the experimental data for the real part of RI of hemoglobin solutions differ for measurements done by alternative techniques (see Table 4), it is important for researchers to use a specific tool, such as the Kramers–Kronig relations, to analyze experimental results for discrete wavelengths and to derive the RI real part theoretically from the measurements of its imaginary part.4,24,49,50,53 In addition to providing quantification of the real part of the RI of hemoglobin at selected wavelengths, where no direct measurements are available, they are independent of hemoglobin concentration and thus can augment the model functions for the RI found by alternative methods.4 Such analysis was done early in Ref. 4 for the measurements of the real part of the RI of hemoglobin solutions at eight discrete wavelengths from 400 to 700 nm, and we received encouraging results. In this work, measurements were done in a wider wavelength range from 480 to 1550 nm at 11 discrete wavelengths, which will allow us to make a more precise Kramers–Kronig analysis, results of which we are planning to publish in the near future.

5.

Conclusion

The RI of hemoglobin solutions has been measured for visible and NIR ranges using a commercially available multiwavelength Atago refractometer. Data were approximated by the Sellmeier formula with a high accuracy in a whole wavelength range. The absolute value of the initial index of refraction n0 and the specific refraction increment dn/dC on hemoglobin concentration C for room temperature at 23°C were derived from these measurements for each wavelength from 480 to 1550 nm. The data obtained are in good agreement with available data in the literature and supplementary to already measured values as done for new wavelengths, which allowed for evaluation of the specific refraction increment dn/dC in a wide spectral range.

Disclosures

The authors have no relevant financial interests in this article and no potential conflicts of interest to disclose.

Acknowledgments

The authors appreciate support from the Tomsk State University Competitiveness Improvement Programme, from the 5 top 100 Russian Academic Excellence Project at the Immanuel Kant Baltic Federal University (ENL), grants of the RFBR 17-02-00358 and the MES RF 17.1223.2017/AP (VVT).

References

1. 

J. J. J. Dirckx, L. C. Kuypers and W. F. Decraemer, “Refractive index of tissue measured with confocalmicroscopy,” J. Biomed. Opt., 10 (4), 044014 (2005). https://doi.org/10.1117/1.1993487 JBOPFO 1083-3668 Google Scholar

2. 

T. L. Troy, D. L. Page and E. M. Sevick-Muraca, “Optical properties of normal and diseased breast tissues: prognosis for optical mammography,” J. Biomed. Opt., 1 (3), 342 –355 (1996). https://doi.org/10.1117/12.239905 JBOPFO 1083-3668 Google Scholar

3. 

A. N. Bashkatov, E. A. Genina and V. V. Tuchin, “Optical properties of skin, subcutaneous, and muscle tissues: a review,” J. Innovative Opt. Health Sci., 4 (1), 9 –38 (2011). https://doi.org/10.1142/S1793545811001319 Google Scholar

4. 

O. Sydoruk et al., “Refractive index of solutions of human hemoglobin from the near-infrared to the ultraviolet range: Kramers-Kronig analysis,” J. Biomed. Opt., 17 (11), 115002 (2012). https://doi.org/10.1117/1.JBO.17.11.115002 JBOPFO 1083-3668 Google Scholar

5. 

L. V. Wang and H. I. Hu, Biomedical Optics: Principles and Imaging, John Wiley and Sons, Hoboken, New Jersey (2007). Google Scholar

6. 

V. V. Tuchin, Tissue Optics: Light Scattering Methods and Instruments for Medical Diagnostics, 988 3rd ed.SPIE Press, Bellingham, Washington (2015). Google Scholar

7. 

C. Meinke, M. Friebel, J. Helfmann, “Optical properties of flowing blood cells,” Advanced Optical Flow Cytometry: Methods and Disease Diagnoses, 95 –132 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim (2011). Google Scholar

8. 

M. Meinke et al., “Optical properties of platelets and blood plasma and their influence on the optical behavior of whole blood in the visible to near infrared wavelength range,” J. Biomed. Opt., 12 014024 (2007). https://doi.org/10.1117/1.2435177 JBOPFO 1083-3668 Google Scholar

9. 

O. S. Zhernovaya, V. V. Tuchin and I. V. Meglinski, “Monitoring of blood proteins glycation by refractive index and spectral measurements,” Laser Phys. Lett., 5 (6), 460 –464 (2008). https://doi.org/10.1002/lapl.200810007 1612-2011 Google Scholar

10. 

P. Giannios et al., “Complex refractive index of normal and malignant human colorectal tissue in the visible and near-infrared,” J. Biophotonics, 10 (2), 303 –310 (2017). https://doi.org/10.1002/jbio.201600001 Google Scholar

11. 

S. Carvalho et al., “Wavelength dependence of the refractive index of human colorectal tissues: comparison between healthy mucosa and cancer,” J. Biomed. Photonics Eng., 2 (4), 040307 (2016). https://doi.org/10.18287/JBPE16.02.040307 Google Scholar

12. 

G. Popescu, Quantitative Phase Imaging of Cells and Tissues, McGraw-Hill, New York (2011). Google Scholar

13. 

H. Majeed et al., “Quantitative phase imaging for medical diagnosis,” J. Biophotonics, 10 (2), 177 –205 (2017). https://doi.org/10.1002/jbio.201600113 Google Scholar

14. 

M. Shan, M. E. Kandel and G. Popescu, “Refractive index variance of cells and tissues measured by quantitative phase imaging,” Opt. Express, 25 (2), 1573 –1581 (2017). https://doi.org/10.1364/OE.25.001573 OPEXFF 1094-4087 Google Scholar

15. 

M. E. Kandel et al., “Label-free tissue scanner for colorectal cancer screening,” J. Biomed. Opt., 22 (6), 066016 (2017). https://doi.org/10.1117/1.JBO.22.6.066016 JBOPFO 1083-3668 Google Scholar

16. 

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

17. 

T. H. Nguyen et al., “Automatic Gleason grading of prostate cancer using quantitative phase imaging and machine learning,” J. Biomed. Opt., 22 036015 (2017). https://doi.org/10.1117/1.JBO.22.3.036015 JBOPFO 1083-3668 Google Scholar

18. 

H. Majeed et al., “Quantifying collagen fiber orientation in breast cancer using quantitative phase imaging,” J. Biomed. Opt., 22 046004 (2017). https://doi.org/10.1117/1.JBO.22.4.046004 JBOPFO 1083-3668 Google Scholar

19. 

M. Lee et al., “Label-free optical quantification of structural alterations in Alzheimer’s disease,” Sci. Rep., 6 31034 (2016). https://doi.org/10.1038/srep31034 Google Scholar

20. 

A. Greenbaum et al., “Wide-field computational imaging of pathology slides using lens-free on-chip microscopy,” Sci. Transl. Med., 6 267ra175 (2014). STMCBQ 1946-6234 Google Scholar

21. 

M. V. Volkenshtein, Molecular Optics, 96 Gosteskhizdat, Moscow (1951). Google Scholar

22. 

G. M. Hale and M. R. Querry, “Optical constants of water in the 200-nm to 200-μm wavelength region,” Appl. Opt., 12 (3), 555 –563 (1973). https://doi.org/10.1364/AO.12.000555 APOPAI 0003-6935 Google Scholar

23. 

D. Segelstein, “The complex refractive index of water,” Department of Physics, University of Missouri, (1981). Google Scholar

24. 

L. X. Cundin and W. P. Roach, “Kramers-Kronig analysis of biological skin,” (2010). Google Scholar

25. 

K. Lee et al., “Measurements of complex refractive indices of photoactive yellow protein,” (2015). Google Scholar

26. 

Z. Deng et al., “Determination of continuous complex refractive index dispersion of biotissue based on internal reflection,” J. Biomed. Opt., 21 (1), 015003 (2016). https://doi.org/10.1117/1.JBO.21.1.015003 JBOPFO 1083-3668 Google Scholar

27. 

F. P. Bolin et al., “Refractive index of some mammalian tissues using a fiber optic cladding method,” Appl. Opt., 28 (12), 2297 –2303 (1989). https://doi.org/10.1364/AO.28.002297 APOPAI 0003-6935 Google Scholar

28. 

M. Daimon and A. Masumura, “Measurement of the refractive index of distilled water from the near-infrared region to the ultraviolet region,” Appl. Opt., 46 (18), 3811 –3820 (2007). https://doi.org/10.1364/AO.46.003811 APOPAI 0003-6935 Google Scholar

29. 

D. K. Sardar and L. B. Levy, “Optical properties of whole blood,” Lasers Med. Sci., 13 (2), 106 –111 (1998). https://doi.org/10.1007/s101030050062 Google Scholar

30. 

G. J. Tearney et al., “Determination of the refractive index of highly scattering human tissue by optical coherence tomography,” Opt. Lett., 20 2258 –2260 (1995). https://doi.org/10.1364/OL.20.002258 OPLEDP 0146-9592 Google Scholar

31. 

A. Knüttel, S. Bonev and W. Knaak, “New method for evaluation of in vivo scattering and refractive index properties obtained with optical coherence tomography,” J. Biomed. Opt., 9 (2), 265 –273 (2004). https://doi.org/10.1117/1.1647544 JBOPFO 1083-3668 Google Scholar

32. 

H.-C. Cheng and Y.-C. Liu, “Simultaneous measurement of group refractive index and thickness of optical samples using optical coherence tomography,” Appl. Opt., 49 790 –797 (2010). https://doi.org/10.1364/AO.49.000790 APOPAI 0003-6935 Google Scholar

33. 

I. Y. Yanina, N. A. Trunina and V. V. Tuchin, “Photoinduced cell morphology alterations quantified within adipose tissues by spectral optical coherence tomography,” J. Biomed. Opt., 18 (11), 111407 (2013). https://doi.org/10.1117/1.JBO.18.11.111407 JBOPFO 1083-3668 Google Scholar

34. 

F. E. Robles, S. Chowdhury and A. Wax, “Assessing hemoglobin concentration using spectroscopic optical coherence tomography for feasibility of tissue diagnostics,” Biomed. Opt. Express, 1 (1), 310 –317 (2010). https://doi.org/10.1364/BOE.1.000310 BOEICL 2156-7085 Google Scholar

35. 

J. H. Jung et al., “Hyperspectral optical diffraction tomography,” Opt. Express, 24 (3), 2006 –2012 (2016). https://doi.org/10.1364/OE.24.002006 OPEXFF 1094-4087 Google Scholar

36. 

F. E. Robles et al., “Molecular imaging true-colour spectroscopic optical coherence tomography,” Nat. Photonics, 5 (12), 744 –747 (2011). https://doi.org/10.1038/nphoton.2011.257 NPAHBY 1749-4885 Google Scholar

37. 

Y. L. Jin et al., “Refractive index measurement for biomaterial samples by total internal reflection,” Phys. Med. Biol., 51 (20), N371 –N379 (2006). https://doi.org/10.1088/0031-9155/51/20/N02 PHMBA7 0031-9155 Google Scholar

38. 

A. García-Valenzuela and H. Contreras-Tello, “Optical model enabling the use of Abbe-type refractometers on turbid suspensions,” Opt. Lett., 38 (5), 775 –777 (2013). https://doi.org/10.1364/OL.38.000775 OPLEDP 0146-9592 Google Scholar

39. 

H. Contreras-Tello and A. García-Valenzuela, “Refractive index measurement of turbid media by transmission of backscattered light near the critical angle,” Appl. Opt., 53 (21), 4768 –4778 (2014). https://doi.org/10.1364/AO.53.004768 APOPAI 0003-6935 Google Scholar

40. 

N. Lue and G. Popescu, “Live cell refractometry using microfluidic devices,” Opt. Lett., 31 (18), 2759 –2761 (2006). https://doi.org/10.1364/OL.31.002759 OPLEDP 0146-9592 Google Scholar

41. 

Y. K. Park et al., “Spectroscopic phase microscopy for quantifying hemoglobin concentrations in intact red blood cells,” Opt. Lett., 34 (23), 3668 –3670 (2009). https://doi.org/10.1364/OL.34.003668 OPLEDP 0146-9592 Google Scholar

42. 

F. E. Robles, L. L. Satterwhite and A. Wax, “Non-linear phase dispersion spectroscopy,” Opt. Lett., 36 (23), 4665 –4667 (2011). https://doi.org/10.1364/OL.36.004665 OPLEDP 0146-9592 Google Scholar

43. 

J.-H. Jung, J. Jang and Y. K. Park, “Spectro-refractometry of individual microscopic objects using swept-source quantitative phase imaging,” Anal. Chem., 85 (21), 10519 –10525 (2013). https://doi.org/10.1021/ac402521u ANCHAM 0003-2700 Google Scholar

44. 

R. Barer, “Refractometry and interferometry of living cells,” J. Opt. Soc. Am., 47 (6), 545 –556 (1957). https://doi.org/10.1364/JOSA.47.000545 JOSAAH 0030-3941 Google Scholar

45. 

D. J. Faber et al., “Oxygen saturation-dependent absorption and scattering of blood,” Phys. Rev. Lett., 93 (2), 028102 (2004). https://doi.org/10.1103/PhysRevLett.93.028102 PRLTAO 0031-9007 Google Scholar

46. 

O. Zhernovaya et al., “The refractive index of human hemoglobin in the visible range,” Phys. Med. Biol., 56 (13), 4013 –4021 (2011). https://doi.org/10.1088/0031-9155/56/13/017 PHMBA7 0031-9155 Google Scholar

47. 

J. Wang et al., “Measurement of the refractive index of hemoglobin solutions for a continuous spectral region,” Biomed. Opt. Express, 6 (7), 2536 –2541 (2015). https://doi.org/10.1364/BOE.6.002536 Google Scholar

48. 

M. Yahya and M. Z. Saghir, “Empirical modelling to predict the refractive index of human blood,” Phys. Med. Biol., 61 1405 –1415 (2016). https://doi.org/10.1088/0031-9155/61/4/1405 PHMBA7 0031-9155 Google Scholar

49. 

S. F. Shumilina, “Dispersion of real and imaginary part of the complex refractive index of hemoglobin in the range 450 to 820 nm,” Bullet. Beloruss. SSR Acad. Sci., 1 79 –84 (1984). Google Scholar

50. 

J. Gienger et al., “Determining the refractive index of human hemoglobin solutions by Kramers–Kronig relations with an improved absorption model,” Appl. Opt., 55 (31), 8951 –8961 (2016). https://doi.org/10.1364/AO.55.008951 APOPAI 0003-6935 Google Scholar

51. 

G. V. Maksimov et al., “Role of viscosity and permeability of the erythrocyte plasma membrane in changes in oxygen-binding properties of hemoglobin during diabetes mellitus,” Bull. Exp. Biol. Med., 140 (5), 510 –513 (2005). https://doi.org/10.1007/s10517-006-0010-x BEXBAN 0007-4888 Google Scholar

52. 

J. Singh, Optical Properties of Condensed Matter and Applications, Wiley, Chichester (2006). Google Scholar

53. 

S. A. Prahl, “Optical absorption of hemoglobin,” (2018) http://omlc.ogi.edu/spectra/hemoglobin/index.html March ). 2018). Google Scholar

54. 

M. Andersen and L. Painter, “Dispersion equation and polarizability of bovine serum albumin from measurements of refractive indices,” Biopolymers, 13 1261 –1267 (1974). https://doi.org/10.1002/(ISSN)1097-0282 BIPMAA 0006-3525 Google Scholar

55. 

M. Friebel and M. Meinke, “Model function to calculate the refractive index of native hemoglobin in the wavelength range of 250 to 1100 nm dependent on concentration,” Appl. Opt., 45 (12), 2838 –2842 (2006). https://doi.org/10.1364/AO.45.002838 APOPAI 0003-6935 Google Scholar

56. 

M. Friebel, “Determination of the complex refractive index of highly concentrated hemoglobin solutions using transmittance and reflectance measurements,” J. Biomed. Opt., 10 (6), 064019 (2005). https://doi.org/10.1117/1.2138027 JBOPFO 1083-3668 Google Scholar

Biography

Ekaterina N. Lazareva is an engineer at the Department of Optics and Biophotonics of Saratov State University. Her research interests are in optical properties of biological tissues, especially of blood (its components) and adipose tissue, refractometry and spectroscopy of tissues, and tissue optical clearing.

Valery V. Tuchin is a professor and head of optics and biophotonics at Saratov State University (National Research University of Russia) and several other universities. His research interests include tissue optics, laser medicine, tissue optical clearing, and nanobiophotonics. He is a fellow of SPIE and OSA, has been awarded Honored Science Worker of the Russia, Honored Professor of Saratov University, SPIE Educator Award, FiDiPro, Finland, Chime Bell Prize of Hubei Province, China, NanQiang Life Science Series Lectures Award of Xiamen University, China, and the Joseph W. Goodman Book Writing Award (OSA/SPIE).

© 2018 Society of Photo-Optical Instrumentation Engineers (SPIE) 1083-3668/2018/$25.00 © 2018 SPIE
Ekaterina N. Lazareva and Valery V. Tuchin "Measurement of refractive index of hemoglobin in the visible/NIR spectral range," Journal of Biomedical Optics 23(3), 035004 (15 March 2018). https://doi.org/10.1117/1.JBO.23.3.035004
Received: 21 November 2017; Accepted: 20 February 2018; Published: 15 March 2018
Lens.org Logo
CITATIONS
Cited by 63 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Blood

Spectroscopy

Refractive index

Near infrared

Temperature metrology

Absorption

Microscopy

Back to Top