Next Article in Journal
High Dynamic Range Imaging with TDC-Based CMOS SPAD Arrays
Next Article in Special Issue
A Low-Cost Phase-OTDR System for Structural Health Monitoring: Design and Instrumentation
Previous Article in Journal
Multidirectional Pin-on-Disk Testing Device to Evaluate the Cross-shear Effect on the Wear of Biocompatible Materials
Previous Article in Special Issue
A Dual-Wavelength Pulsed Laser Processing Platform for a-Si Thin Film Crystallization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Technical Note

Critical-Angle Differential Refractometry of Lossy Media: A Theoretical Study and Practical Design Issues

by
Spyridon Koutsoumpos
,
Panagiotis Giannios
,
Dimos Triantis
and
Konstantinos Moutzouris
*
Electronic Devices & Materials Lab, Department of Electrical & Electronic Engineering, University of West Attica, 12243 Athens, Greece
*
Author to whom correspondence should be addressed.
Current address: Barcelona Institute of Science and Technology, Institute for Research in Biomedicine, IRB Barcelona 08028, Spain.
Instruments 2019, 3(3), 36; https://doi.org/10.3390/instruments3030036
Submission received: 3 June 2019 / Revised: 22 July 2019 / Accepted: 30 July 2019 / Published: 31 July 2019
(This article belongs to the Special Issue Photonic Devices Instrumentation and Applications)

Abstract

:
At a critical angle of incidence, Fresnel reflectance at an interface between a front transparent and a rear lossy medium exhibits sensitive dependencies on the complex refractive index of the latter. This effect facilitates the design of optical sensors exploiting single (or multiple) reflections inside a prism (or a parallel plate). We determine an empirical framework that captures performance specifications of this sensing scheme, including sensitivity, detection limit, range of linearity and—what we define here as—angular acceptance bandwidth. Subsequently, we develop an optimization protocol that accounts for all relevant optical or geometrical variables and that can be utilized in any application.

1. Introduction

Refractive index sensors are intensely investigated for numerous biomedical [1,2,3], chemical [4,5] and industrial [6,7] applications. An indicative yet far from exhaustive list of sensing mechanisms relies on plasmonic [8,9,10,11], photonic crystal [12,13,14,15], micro-cavity [16,17,18,19], optical fiber [20,21,22,23] and wave-guide [24,25,26,27] configurations. Associated with Fresnel reflectance properties at planar interfaces, differential refractometry offers an alternative path to sensing refractive index changes, by exploitation of interference [28], deflection [29] or (more relevant to the present work) critical-angle [30,31,32,33,34,35] effects. Today, differential refractometry is not only a standard analytical tool that operates routinely in many laboratories, but also infiltrates emerging optofluidic and lab-on-chip technologies [36,37,38].
In critical-angle differential refractometry (CADR), a front transparent medium (commonly, a prism) is interfaced with a sample which, typically, is also assumed to be lossless. The underlying principle of operation is simple: provided that the front medium is optically denser than the sample, there exists a sharp transition from total internal reflection (TIR) to partial internal reflection, taking place at a critical angle which corresponds to the location of an abrupt discontinuity in the derivative of reflectance with respect to incidence angle. Operating the sensing interface at the transition point leads to the generation of an intensity readout signal, as soon as the TIR condition is disturbed by refractive index fluctuations.
With non-transparent samples, CADR interpretation is less straightforward. To begin with, the refractive index of the rear medium becomes a complex quantity, the imaginary part of which incorporates absorption or scattering effects. Furthermore, reflectance never reaches unity (except for the limiting case of incidence at 90° with respect to the surface normal) and the transition from “attenuated” total internal reflection (ATIR) to partial internal reflection is gradual. As a result, the reflectance derivative with respect to incidence angle, now peaking to a finite value, is no longer the proper quantity to conceptualize the sensing principle; this purpose is better served by reflectance derivatives with respect to the real and imaginary index of the sample. These are negative quantities exhibiting local extrema at the vicinity of the transition from ATIR to partial internal reflection, albeit at slightly different “critical” angles. In general, these extrema are stronger for p- than for s-polarization, an observation that indicates the preferential wave orientation for CADR sensing.
In this work, we attempt a theoretical study of CADR with lossy media, accounting for (i) the standard prism configuration, (ii) an alternative geometry that exploits multiple reflections inside a parallel plate. Both schemes are, in principle, compatible with optofluidic technologies and static or real-time monitoring applications. We untangle the perplex dependencies of the sensor’s specifications on various optical and geometrical parameters, revealing among other facts the dominant role of the sample’s loss. In doing so, we introduce the concept of the “angular acceptance bandwidth” which trades-off with sensor’s sensitivity and helps clarify several complexities in terms of light coupling and detection. Our results provide a universal roadmap for rapid performance evaluation and optimization of CADR devices, which might be essential for pushing the technique’s detection limit from the current standard ( 100   μ RIU ) down to the current state-of-the-art ( 1   μ RIU for CADR [31], as well as for all noninterferometric methods), or even lower.

2. Theoretical Background

The isosceles triangular prism and parallelogram plate configurations under investigation are depicted in Figure 1. The sample is interfaced with the base of the prism, or equivalently, the top side of the plate, the bottom of which is high-reflection-coated. Coming from air, light hits the input facet at an external angle of incidence θ and reaches the output facet at an angle ϕ , after N = 1 (or N > 1 ) reflections at the sensing interface of the prism (plate), respectively; each one of these reflections is at a constant angle θ 1 . The input and output coupling geometries are determined by the cut angle α , which in Figure 1 is taken α < π / 2 ; however, the upcoming formalism remains valid also for α π / 2 . Then, θ , ϕ and θ 1 (which are trivially defined with respect to the corresponding normal), as well as plate’s length L and thickness d (which is assumed to be much larger than the light wavelength, so as to avoid interference effects), relate via:
θ 1 = α ϕ , ϕ = a r c s i n n a i r n o · s i n θ , L 2 d · N · t a n ( θ 1 )
where n a i r ( 1 ) and n o are the real indices of air and the transparent prism (or plate), respectively.
Light transfer in the proposed layouts can be accurately simulated via standard Fresnel theory. Let us begin by assuming that a reference medium with a complex refractive index ( n = n r i n i ) is first put to the test. The “reference” prim/plate transmittance T p , defined as the ratio of the output light intensity I r e f over incident light intensity I i n , is:
T p ( θ ) = I r e f I i n = 1 R 1 / 2 ( θ ) · R 2 / 3 ( θ 1 ) N · 1 R 2 / 1 ( ϕ )
where R k / j is the reflectance at the interface between medium k and medium j ( k , j = 1 , 2 , 3 with 1 = a i r , 2 = p r i s m / p l a t e , 3 = s a m p l e ); Reflectance at the metal-coated surface of the plate is taken as unity; Equation (2) is valid for the prism, as well as the plate, by proper substitution of the respective N value. In turn, R k / j is the squared modulus of the corresponding Fresnel amplitude coefficient, which—for the preferred p-polarization—reads:
r k / j ( ) = c o s ( ) · ( n j / n k ) 2 ( n j / n k ) 2 s i n 2 ( ) c o s ( ) · ( n j / n k ) 2 + ( n j / n k ) 2 s i n 2 ( )
where = θ , θ 1 , ϕ for ( k / j = 1 / 2 , 2 / 3 , 2 / 1 ) and n k , n j are the refractive indices of the respective media.
Then, suppose that the reference medium is replaced by a corrupted sample, which exhibits a complex index
n = ( n r + δ n r ) i · ( n i + δ n i ) = n + ( δ n r i · δ n i )
that is, equal to the refractive index n of the reference medium, with an added perturbation term that comprises a real ( δ n r ) and an imaginary ( δ n i ) component. At this point, we may appropriately define a total index perturbation parameter δ via:
δ = δ n r + δ n i .
This index perturbation term shifts the plate transmittance (and output intensity) to new values T (and I) that can be straightforwardly calculated by substituting n for n in Equation (2). Assuming that the incident light intensity I i n remains constant, the relative intensity change Δ I / I between measurements of the corrupted and reference sample is:
Δ I I = I I r e f I = T T p T
Figure 2 depicts the reference transmittance T p , along with the relative intensity change Δ I / I , versus external angle of incidence θ ; specifics on parameters used for calculations are given in the figure’s label. (Note that the assumption of unit reflectance at the metal-coated surface may lead to somewhat overestimated T p values, but does not affect the relative intensity change.) Apparently, Δ I / I exhibits local extremum at a critical external angle θ = θ c , which shifts only slightly, when a perturbation in the real index alone ( δ = δ n r ) is replaced by an equal-sized perturbation in the imaginary index ( δ = δ n i ). Interestingly, these two cases are distinguishable by the asymmetric shape of the relative intensity change curve; indeed, when δ = δ n r ( δ = δ n i ), Δ I / I decays faster towards smaller (larger) incidence angles, respectively. These observations reflect similar properties of the reflectance derivatives with respect to n r , n i and may be relevant for applications requiring sensor’s selectivity.

3. Signal, Sensitivity and Acceptance Bandwidth

From a practical standpoint, the most important features of the relative intensity change curve are (i) its extreme value at the critical angle Δ I / I c , thereon understood as the sensor’s output signal, (ii) its angular bandwidth A B W , defined as the full-width-at-half-maximum and thereon understood as the sensor’s acceptance bandwidth. Due to the entangled dependencies on multiple variables, attempting to evaluate the sensor’s signal and acceptance bandwidth by generating successive plots like those presented in Figure 2, is an elaborate, time-demanding and somewhat confusing task. Fortunately, we found out that it is possible to produce a set of empirical equations that can handily trace both Δ I / I c and A B W (more details on the heuristic model-building process are given in Appendix A). These equations are:
Δ I I c = S · δ
S F n r / n o · n i n o a o · N n o · δ 1 ,
A B W 10 5 · d θ 1 d θ | θ = θ c 1 · G n r / n o · n i n o · δ 2 ,
where S is sensor’s sensitivity, A B W is given in degrees and:
δ 1 = 1 1 2 δ n r δ + 1 2 δ n r δ 2 , δ 2 = 1 δ n r δ + δ n r δ 2
are factors acquiring a maximum value of 1 when δ = δ n r + δ n i = δ n r or δ = δ n r + δ n i = δ n i , and a minimum value when δ n r = δ n i = δ / 2 ,
F n r / n o = a 1 · e x p [ a 2 n r n o 1 ] + a 3 1 a 4 n r n o 1 ,
G n r / n o = a 5 + a 6 · 1 a 7 n r n o 1 / 2 ,
and
d θ 1 d θ | θ = θ c = c o s θ c · n a i r n o 2 s i n 2 θ c 1 / 2 ,
with: a o = 0.49458 , a 1 = 1.84 , a 2 = 0.695658 , a 3 = 0.267344 , a 4 = 0.990367 , a 5 = 0.000717133 , a 6 = 0.00259023 , a 7 = 1.00072 .
Equations (7)–(9) reproduce the sensor’s signal, sensitivity and acceptance bandwidth with a typical accuracy of second significant digit, which is sufficient for evaluation and optimization purposes, as long as the respective variables remain within the following ranges:
δ < n i , N < 50 , 0.5 < n r n o < 1 , 10 2 < n i n o < 10 8 ,
The information contained in these empirical equations is graphically exposed in Figure 3. Therein, it is clear that Δ I / I c depends explicitly on δ , δ n r / δ , N / n o , n r / n o and n i / n o . These dependencies are linear with respect to δ and N / n o , but nonlinear with respect to the remaining three variables, such that the signal maximizes when: δ n r / δ = 0 o r 1 , n r / n o 1 and, more rapidly, as n i / n o reduces. It is noted that sensitivity S is by definition a direct measure of the output signal—after normalization by the input stimuli δ —and thus, it behaves likewise.
Comparatively, A B W exhibits nonlinear dependencies on δ n r / δ , n r / n o and scales linearly with n i / n o ; however, it is not affected by δ or N. Contrary to Δ I / I c , which is geometry-independent, A B W varies also with the critical angle θ c . This nonlinear dependence emerges from the derivative d θ 1 / d θ | θ = θ c , which translates the internal angular bandwidth (in terms of θ 1 ) into an external one (in terms of θ ). Note that the critical angle θ c is tunable by means of the plate’s angle α , as is indicated by Equation (1). Derivative d θ 1 / d θ | θ = θ c introduces also a relatively weak dependence on n a i r / n o , which can be neglected at first instance and, thus, is not included in Figure 3. To sum up, angular acceptance bandwidth maximizes when: δ n r / δ = 0 o r 1 , n r / n o 1 , θ c 90 ° and, more rapidly, as n i / n o increases.
Remarkably, the main regulator of the sensor’s sensitivity (or equivalently, signal) and acceptance bandwidth is the sample’s imaginary index, which introduces a trade-off between these quantities. Indeed, at the limit of transparency ( n i = 0 ) Equations (8) and (9) return an infinite S and a null A B W ; this result conforms with the standard CARD interpretation, conveying features of the reflectance derivative with respect to incidence angle, at the point of discontinuity. Inversely, as loss increases, sensitivity tends to zero and acceptance bandwidth grows larger.
The only means available for simultaneous increase in S and A B W , which is an obvious-seeming operational advantage, is to stretch parameters { n r / n o , N , θ c } close to their limiting values { 1 , 1 , 90 ° } . Although appealing in principle, these limits should be avoided in practise, because they diminish the overall prism/plate transmittance, an effect that intensifies with increasing loss and complicates output light detection. There exists two more reasons to avoid N 1 . First, large N values deteriorate sensor’s performance if the plate is not perfectly parallel, a repercussion that is otherwise minor, considering the sub- μ r a d tolerances that are nowadays available with state-of-the-art plane-parallel optics. Second, N 1 corresponds to large plate lengths, compromising the device’s compactness and increasing sample volume demands.

4. Detection Limit Considerations

The trade-off between sensitivity and angular acceptance bandwidth raises some interesting questions; does an optimal value for n i and by extension, a preferred wavelength of operation, exist? If so, how can it be determined? The compromise solution to these problems emerges via detection limit ( D L ) considerations.
The detection limit, i.e., the minimum index perturbation that can be reliably sensed, is an instrument, and application-specific quantity that can be estimated, provided that significant sources of noise are distinguishable. For example, let us assume that non-negligible noise contributions originate from (i) the detection stage ( σ d e t ) , (ii) power fluctuations of the laser source ( σ l a s e r ), (iii) random fluctuations in the real index of the sample, caused by temperature instabilities Δ T and the nonzero spectral bandwidth Δ λ of the laser source, which introduce noise variance components σ T = S · [ d n r / d T ] · Δ T and σ λ = S · [ d n r / d λ ] · Δ λ , respectively. These assumptions are reasonable, as long as the thermooptic and dispersion coefficients for the real index of the sample (i.e., d n r / d T and d n r / d λ ) are much larger than the respective coefficients for n i , n o , and n a i r .
Then, we may adopt standard convention D L = 3 σ / S , where σ σ l a s e r 2 + σ T 2 + σ λ 2 + σ d e t 2 is the total noise variance, so as to deduce the following useful approximation:
D L 3 σ l a s e r 2 + S d n r d T Δ T 2 + S d n r d λ Δ λ 2 + σ d e t 2 / S .
It is appropriate at this point to parenthetically note that Equation (15) combined with the “rule of the thumb” articulated in Equation (14), specify the range of sensor’s linear response, which is the last remaining major specification that depends also on sample’s loss:
L i n e a r r a n g e : D L < δ < n i .
Under the assumptions validating Equation (15), the detection limit decreases monotonically with increasing sensitivity, reaching its minimum theoretical value D L m i n at the transparency limit (that is, where S ):
D L m i n = lim S D L = 3 d n r d T Δ T 2 + d n r d λ Δ λ 2
However, reduction in the detection limit by means of sensitivity increase is meaningful only up to the point that S = S o p t , where
S o p t = σ l a s e r 2 + σ d e t 2 d n r / d T Δ T 2 + d n r / d λ Δ λ 2 .
Indeed, when S = S o p t , the detection limit reaches its “optimum” value D L o p t that is equal to:
D L o p t = 2 D L m i n = 3 2 d n r d T Δ T 2 + d n r d λ Δ λ 2 .
Further increase in sensitivity towards infinity (by decreasing n i towards the transparency limit) has a negligible effect on D L , while drastically reducing the acceptance bandwidth. Besides, higher sensitivities also compromise sensor’s range of linearity, as may be seen by inspection of Equation (15). Therefore, operating the sensing interface with S > S o p t is not only unnecessary but also detrimental.

5. Spatially-Unresolved vs. Spatially-Resolved Detection

Setting S = S o p t by choosing the proper value for n i automatically locks angular acceptance bandwidth to its own respective value, which is easily determinable, as well as decisive for the design of the detection stage. The latter can be based either on a spatially-unresolved, or on a spatially-resolved, scheme.
Accurate spatially-unresolved detection (SUD) by use of a standard photodiode is the most straightforward option; it requires that the input light is stably coupled into the prism/plate at the critical angle θ c , with its entire optical power contained within an angular spread that does not exceed A B W . Assuming an input beam with high pointing-stability, waist radius w o and quality factor M 2 , the full-angle divergence θ d at the diffraction-limit is:
θ d = 2 M 2 λ π w o
and the implied condition for SUD detection reads:
S U D c o n d i t i o n : A B W θ d 1 .
As A B W becomes narrower (i.e., S o p t increases), the SUD condition necessitates beams with large waists. When Equation (21) is not fulfilled, the signal hides inside a fraction of the beam profile and, hence, spatially-resolved detection (SRD) by use of a diode array should be opted. Accurate SRD requires that the “active area” of the output beam profile (i.e., the fraction containing the sensing information) should be larger than the surface of at least one pixel of the diode array, that is:
S R D c o n d i t i o n : d p i x e l 2 w < A B W θ d ,
where d p i x e l is the pixel’s diameter and w is the beam radius at the detector.
When either of these conditions is fulfilled, detection noise may become negligible and thus, it is eliminated from Equation (15) and (18). Indeed, removing measurement complications associated with the finite acceptance bandwidth can reduce σ d e t down to levels that relate only to fundamental noise mechanism, such as thermal and quantum noise. Then, we may reasonably assume that (i) the noise-equivalent-power at the photodiode will be in the sub- n W range, (ii) output light power reaching the detector will far exceed 1 μ W . These assumptions confirm the estimate that σ d e t 10 3 , which is indeed well-below the power noise σ l a s e r of common laser sources.

6. Optimization Protocol

Following the preceding analysis, it is now possible to formulate a simple optimization process that can be always adopted, as long as the optical constants of the reference sample (i.e., n r , n i , d n r / d T , d n r / d λ ) and their wavelength dependence are known. For the sake of simplicity, we may select δ 1 = δ 2 = 1 , since in most practical cases an extraneous stimulus will affect primarily either the real or the imaginary index of the sample. The proposed protocol comprises the following steps:
Step 1: Impose strict temperature regulation and narrow-band laser emission (or spectral filtering), so as to reduce Δ T and Δ λ ; for any given sample, these are the only free variables affecting the detection limit, which can be calculated by use of Equation (19).
Step 2: Compute the optimum sensitivity S o p t via Equation (18); it is at this stage that laser power noise σ l a s e r becomes relevant, while as previously explained, detection noise σ d e t can be considered practically negligible.
Step 3: Select a transparent solid with known index n o > n r that can be shaped into a prism. Initially, avoid close-matching between the refractive indices of the front medium and the sample, so as to ensure substantial total transmittance. For the same reason, the number of reflections is initially set to N = 1 .
Step 4: Given the light extinction spectrum of the sample, identify the laser wavelength that corresponds to the value of n i which, when substituted to Equation (8), allows sensitivity to reach its optimal value S o p t .
Step 5: With all involved variables being determined, calculate the respective angular acceptance bandwidth A B W using Equation (9); to sustain high total transmittance, select initially the external critical angle θ c for grazing incidence.
Step 6: By use of the criteria shown in Equations (21) and (22), evaluate the feasibility of spatially-unresolved and spatially-resolved detection geometries. If found practical, the optimization process is complete and the setup fully drawn accordingly. Otherwise, go on to the next step:
Step 7: Repeat the process so as to increase A B W , by selecting an optically denser transparent medium, a parallel plate with multiple reflections instead of a prism, and oblique incidence, in order to approach the operational limits { n r / n o , N , θ c } { 1 , 1 , 90 ° } . Ensure that the total transmittance is high enough to empower output light detection and the resulting plate length is small enough to comply with application’s needs.

7. Further Discussion and Concluding Remarks

To demonstrate the susceptibility of the sensor’s specifications to the various parameters involved, Table 1 presents results from the application of the proposed protocol to three hypothetical scenarios. Chosen values of variables remain within realistic ranges, being very tolerant in Case 1 and becoming more and more rigid as we move to Cases 2 and 3. Top rows in Table 1 group together noise-related parameters Δ T , Δ λ and σ l a s e r ( σ d e t is considered negligible). The second set accounts for sample’s optical properties, namely n r , n i , d n r / d T , d n r / d λ , while the third set incorporates information regarding the transparent medium and the prism/plate geometry, that is n o , N and θ c . Based on these values, acceptance bandwidth A B W and ratios A B W / θ d are then evaluated. At the bottom of Table 1, corresponding values for the sensitivity S, optimal sensitivity S o p t and detection limit D L are given. In accordance with the protocol’s guidelines, the sample’s imaginary index n i is always chosen such that sensitivity S, quantified via Equation (8), equals its optimal value S o p t which, in its own turn, is determined via Equation (18).
Several points of interests emerge from data in Table 1. Depending on the magnitude of Δ T and Δ λ , the detection limit varies from 600 μ RIU down 30 μ RIU and sub- μ RIU levels, as we move on from Case 1 via Case 2 to Case 3; D L reduction is accompanied by (i) increase in S o p t from 71 to 710 and 7100, (ii) decrease in n i from 3 × 10 3 to 1.3 × 10 4 and 1 × 10 5 , (iii) narrowing of A B W from 1.82 ° to 0.12 ° and 0.03 ° , respectively. Corresponding A B W / θ d values indicate that spatially-unresolved detection is possible with beam waists w o 0.02 mm, 0.37 mm and 1.25 mm, respectively, assuming for the sake of argument an operational wavelength in the 1 μ m range. To meet these specifications, parameters { n r / n o , N , θ c } changed from initial values { 0.75 , 1 , 15 ° } in Case 1, to { 0.83 , 2 , 45 ° } in Case 2 and { 0.89 , 4 , 75 ° } in Case 3; corresponding values for the prism/plate cut angle are α = 55.95 ° , 79.57 ° and 106.77 ° . As a final check, these changes cause no severe cutback to the overall plate transmittance, which always exceeds 50 % , requiring, however, plate lengths as large as L 3 cm in Case 2 and L 11 cm in Case 3, for an assumed plate thickness d = 0.5 cm.
If such plate lengths are inappropriate, parameters { n r / n o , N , θ c } can be kept constant to their initial values (i.e., { 0.75 , 1 , 15 ° } ). The detection limit would then be unaffected, but acceptance bandwidth would be reduced in Cases 2 and 3, such that spatially-unresolved detection would require beam waists w o 2.5 mm and 250 mm, respectively. Therefore, spatially-resolved detection would become a reasonable option in Case 2 and an inevitable choice in Case 3 for which, a maximum pixel diameter of 16 μ m is specified, when plugging into the SRD criterion indicative values w o 1 mm and w 2 mm.
As a further practical demonstration on the relevance of our analysis to real-life applications, Figure 4 depicts the wavelength scaling of S and A B W for two interfaces comprising water as rare medium and (i) calcium fluoride, (ii) standard S F 10 glass, as front media. Water is chosen as an exemplary test medium since it is the main constituent of biofluids/tissues and a highly relevant medium for environmental sensing and quality control purposes. Calculations adopt values for the required optical constants (that is, the wavelength-dependent complex index of water and the wavelength-dependent real indices of C a F 2 and S F 10 , which can be considered transparent throughout the spectral range of interest) from references [39,40,41]. Within the near-infrared band 0.9 μ m– 1.6 μ m, S and A B W can be tuned by approximately two and three orders of magnitude, respectively, for the SF10/water interface and indicative values of N = 1 , θ c = 15 ° . This behavior reflects primarily the absorption profile of water, which exhibits a minimum at 0.9 μ m ( n i 4.5 × 10 7 ) and a maximum at 1.46 μ m ( n i 3.2 × 10 4 ). Further tuning is possible by increasing the values of N, θ c and simultaneously using C a F 2 as front medium, whose index matches better the real index of water, thus corresponding to larger values of n r / n o . Figure 4 reveals the heavy impact of laser wavelength on the sensor’s specifications, an effect that does not apply only to water, but practically extends to all non-black, liquid or solid media that exhibit absorption profiles containing nearly-transparent regions, along with absorption peaks. Therefore, multipurpose critical angle refractometers, which are nowadays commercially available and typically equipped with just one or few single-wavelength lasers, would benefit significantly from exploiting tunable sources of coherent radiation, such as continuous-wave optical parametric oscillators [42].
To conclude, in this technical note we thoroughly examined the sensing properties of transparent/lossy interfaces. An empirical framework was developed to describe performance specifications of corresponding sensing geometries that exploit either a single reflection inside a prism, or multiple reflections inside a parallel plate. The dominant role of sample’s loss was revealed and quantified through a trade-off between sensitivity on one hand and angular acceptance bandwidth on the other. These observations enabled the establishment of a generically applicable optimization’s protocol that is based on two main realizations. First, a noise-dependent optimal value S o p t does exist, so that the quest for ever-increasing sensitivities is not just pointless, but also undesirable. Second, operating the sensing interface with S = S o p t directly locks acceptance bandwidth A B W to a calculable value, which then imposes definable conditions for the spatially-unresolved—or spatially-resolved—detection of the signal. The exemplary application of the proposed optimization protocol to three hypothetical scenarios revealed the critical dependence of sensor’s specifications on the relevant optical and geometrical variables; this fact was further supported by calculations of the—widely varying—sensitivity S and acceptance bandwidth A B W , as a function of wavelength, in the representative cases of the C a F 2 /water and S F 10 /water interfaces. In this sense, the present note clarifies several design issues in the area of optical refractometry, relating to the pump source (e.g., choice of wavelength), the mounting medium (e.g., choice of prism material and geometry), as well as the detection stage (that is, spatially unresolved versus spatially resolved approach). We thus anticipate that our study will serve as a valuable tool for the realization of optimized CADR setups, which will push the limit of detection towards the sub- μ RIU regime.

Author Contributions

Conceptualization, K.M.; methodology, S.K., P.G., D.T. and K.M.; validation, S.K., P.G. and K.M.; formal analysis, S.K., P.G. and K.M.; investigation, S.K., P.G., K.M.; writing–original draft preparation, K.M. and D.T.; writing–review and editing, K.M., D.T., P.G. and S.K.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. On the derivation of the empirical Equations (8) and (9)

Model building with data fitting was the basis for deriving the empirical Equations (8)—equivalently, (7)—and (9), which reproduce the sensitivity—or signal—and the angular acceptance bandwidth, respectively. The heuristic process involved the following steps:
(i). First, we attempted to identify the independent variables of the problem, out of a total number of nine initial variables, namely: { n o , n r , n i , n a i r , δ n r , δ n i , N , θ c , α } . To do so, we generated a large set of relative-intensity-change versus external-angle-of-incidence plots; from these plots, we were able to calculate “exact” values of S = Δ I / I c / δ and A B W for various values of the nine starting variables. Close inspection of data revealed that S is practically independent on { n a i r , θ c , α } , while A B W is practically independent on { N , α } . Inspired by the fact that the refractive indices of different media always enter in ratios in Fresnel equations, we were also able to establish that variables { n o , n r , n i , n a i r , N } are not independent, but can be reduced to a set of four variables, namely { n r / n o , n i / n o , n a i r / n o , N / n o } . This realization was simply based on the observation that any changes in the values of { n o , n r , n i , n a i r , N } do not affect S and A B W , provided that the ratios { n r / n o , n i / n o , n a i r / n o , N / n o } remain constant. Similarly, we noticed that variables { δ n r , δ n i } can be reduced to only one independent variable, which is { δ n r / ( δ n r + δ n i ) } –or equivalently, { δ n r / δ } . We thus concluded that independent variables are: { n r / n o , n i / n o , N / n o , δ n r / δ } for S; { n r / n o , n i / n o , n a i r / n o , δ n r / δ , θ c } for A B W .
(ii). Then, we attempted to capture the dependence of S and A B W on each one of the independent variables, through least-square fitting of generated data. The employed functional forms (namely, exponentials, polynomials and rational functions, as shown in Equations (10)–(12)) are largely arbitrary, in the sense that they were chosen from a large set of different models, according to their simplicity and capability to best-agree with the generated data, rather than as a consequence of an underlying physical law. The only exception to this fully-empirical process accounts for the dependence of A B W on { n a i r / n o , θ c } . In this case, we realized that the angular acceptance bandwidth physically originates from reflections at the interface between the prism (or plate) and the sample, and thus relates fundamentally with the internal angle θ 1 . Throughout the main text, we use the external angle θ instead of θ 1 to define relevant quantities, since this is the actual angle of interest for the end-user. Therefore, we made the reasonable assumption that the derivative [ d θ 1 / d θ ] 1 should also be included as a term in the empirical equation for A B W , in order to “translate” the internal angular bandwidth (in terms of θ 1 ) into an external one (in terms of θ ). It is indeed this derivative that introduces the dependencies of A B W on variables { n a i r / n o , θ c } (see Equation (13)).
(iii). Finally, we generated additional data of S and A B W (other than those used for fitting purposes), in order to establish the quality of the proposed empirical equations—Figure 3 is an exemplary demonstration of this test. We found out that the empirical model reproduces synthetic data with a typical accuracy of two significant digits which, as also claimed in the main text, is more than sufficient for evaluation purposes.

References

  1. Bekmurzayeva, A.; Dukenbayev, K.; Shaimerdenova, M.; Bekniyazov, I.; Ayupova, T.; Sypabekova, M.; Molardi, C.; Tosi, D. Etched Fiber Bragg Grating Biosensor Functionalized with Aptamers for Detection of Thrombin. Sensors 2018, 18, 4298. [Google Scholar] [CrossRef] [PubMed]
  2. Chiavaioli, F.; Gouveia, C.; Jorge, P.; Baldini, F. Towards a Uniform Metrological Assessment of Grating-Based Optical Fiber Sensors: From Refractometers to Biosensors. Biosensors 2017, 7, 23. [Google Scholar] [CrossRef] [PubMed]
  3. Giannios, P.; Toutouzas, K.G.; Matiatou, M.; Stasinos, K.; Konstadoulakis, M.M.; Zografos, G.C.; Moutzouris, K. Visible to near-infrared refractive properties of freshly-excised human-liver tissues: marking hepatic malignancies. Sci. Rep. 2016, 6, 27910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Tan, R.; Yap, S.; Tan, Y.; Tjin, S.; Ibsen, M.; Yong, K.; Lai, W. Functionalized Fiber End Superstructure Fiber Bragg Grating Refractive Index Sensor for Heavy Metal Ion Detection. Sensors 2018, 18, 1821. [Google Scholar] [CrossRef] [PubMed]
  5. Sequeira, F.; Duarte, D.; Bilro, L.; Rudnitskaya, A.; Pesavento, M.; Zeni, L.; Cennamo, N. Refractive Index Sensing with D-Shaped Plastic Optical Fibers for Chemical and Biochemical Applications. Sensors 2016, 16, 2119. [Google Scholar] [CrossRef] [PubMed]
  6. Oelhafen, J.; Mayr, T.; Dorner, F.; Moutzouris, K.; Roths, J.; Drechsler, K. Fiber Optic Measurement System for Fresnel Reflection Sensing: Calibration, Uncertainty, and Exemplary Application in Temperature-Modulated Isothermal Polymer Curing. J. Light. Technol. 2018, 36, 939–945. [Google Scholar] [CrossRef]
  7. Munap, D.H.A.; Bidin, N.; Islam, S.; Abdullah, M.; Marsin, F.M.; Yasin, M. Fiber Optic Displacement Sensor for Industrial Applications. IEEE Sens. J. 2015, 15, 4882–4887. [Google Scholar] [CrossRef]
  8. Klinghammer, S.; Uhlig, T.; Patrovsky, F.; Böhm, M.; Schütt, J.; Pütz, N.; Baraban, L.; Eng, L.M.; Cuniberti, G. Plasmonic Biosensor Based on Vertical Arrays of Gold Nanoantennas. ACS Sens. 2018, 3, 1392–1400. [Google Scholar] [CrossRef] [PubMed]
  9. Lal, S.; Link, S.; Halas, N.J. Nano-optics from sensing to waveguiding. Nat. Photonics 2007, 1, 641–648. [Google Scholar] [CrossRef]
  10. Guerreiro, A.; Santos, D.F.; Baptista, J.M. New Trends in the Simulation of Nanosplasmonic Optical D-Type Fiber Sensors. Sensors 2019, 19, 1772. [Google Scholar] [CrossRef]
  11. Kroh, C.; Wuchrer, R.; Steinke, N.; Guenther, M.; Gerlach, G.; Härtling, T. Hydrogel-Based Plasmonic Sensor Substrate for the Detection of Ethanol. Sensors 2019, 19, 1264. [Google Scholar] [CrossRef] [PubMed]
  12. Xu, P.; Zheng, J.; Zhou, J.; Chen, Y.; Zou, C.; Majumdar, A. Multi-slot photonic crystal cavities for high-sensitivity refractive index sensing. Opt. Express 2019, 27, 3609. [Google Scholar] [CrossRef] [PubMed]
  13. Qiao, Q.; Xia, J.; Lee, C.; Zhou, G. Applications of Photonic Crystal Nanobeam Cavities for Sensing. Micromachines 2018, 9, 541. [Google Scholar] [CrossRef] [PubMed]
  14. Maurya, J.; François, A.; Prajapati, Y. Two-Dimensional Layered Nanomaterial-Based One-Dimensional Photonic Crystal Refractive Index Sensor. Sensors 2018, 18, 857. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, H.; Yan, X.; Li, S.; An, G.; Zhang, X. High Sensitivity Refractive Index Sensor Based on Dual-Core Photonic Crystal Fiber with Hexagonal Lattice. Sensors 2016, 16, 1655. [Google Scholar] [CrossRef] [PubMed]
  16. Tsigaridas, G.N. A study on refractive index sensors based on optical micro-ring resonators. Photonic Sens. 2017, 7, 217–225. [Google Scholar] [CrossRef]
  17. Arnold, S.; Khoshsima, M.; Teraoka, I.; Holler, S.; Vollmer, F. Shift of whispering-gallery modes in microspheres by protein adsorption. Opt. Lett. 2003, 28, 272. [Google Scholar] [CrossRef] [PubMed]
  18. Zhu, H.; White, I.M.; Suter, J.D.; Dale, P.S.; Fan, X. Analysis of biomolecule detection with optofluidic ring resonator sensors. Opt. Express 2007, 15, 9139. [Google Scholar] [CrossRef]
  19. Liu, W.; Yan, J.; Shi, Y. High sensitivity visible light refractive index sensor based on high order mode Si_3N_4 photonic crystal nanobeam cavity. Opt. Express 2017, 25, 31739. [Google Scholar] [CrossRef]
  20. Tsigaridas, G.; Polyzos, D.; Ioannou, A.; Fakis, M.; Persephonis, P. Theoretical and experimental study of refractive index sensors based on etched fiber Bragg gratings. Sens. Actuators A Phys. 2014, 209, 9–15. [Google Scholar] [CrossRef]
  21. Ahsani, V.; Ahmed, F.; Jun, M.; Bradley, C. Tapered Fiber-Optic Mach-Zehnder Interferometer for Ultra-High Sensitivity Measurement of Refractive Index. Sensors 2019, 19, 1652. [Google Scholar] [CrossRef] [PubMed]
  22. Pugliese, D.; Konstantaki, M.; Konidakis, I.; Ceci-Ginistrelli, E.; Boetti, N.G.; Milanese, D.; Pissadakis, S. Bioresorbable optical fiber Bragg gratings. Opt. Lett. 2018, 43, 671. [Google Scholar] [CrossRef] [PubMed]
  23. Apriyanto, H.; Ravet, G.; Bernal, O.D.; Cattoen, M.; Seat, H.C.; Chavagnac, V.; Surre, F.; Sharp, J.H. Comprehensive Modeling of Multimode Fiber Sensors for Refractive Index Measurement and Experimental Validation. Sci. Rep. 2018, 8, 5912. [Google Scholar] [CrossRef]
  24. Wang, M.; Zhang, M.; Wang, Y.; Zhao, R.; Yan, S. Fano Resonance in an Asymmetric MIM Waveguide Structure and Its Application in a Refractive Index Nanosensor. Sensors 2019, 19, 791. [Google Scholar] [CrossRef] [PubMed]
  25. Guo, Z.; Wen, K.; Hu, Q.; Lai, W.; Lin, J.; Fang, Y. Plasmonic Multichannel Refractive Index Sensor Based on Subwavelength Tangent-Ring Metal–Insulator–Metal Waveguide. Sensors 2018, 18, 1348. [Google Scholar] [CrossRef] [PubMed]
  26. Hong, Y.S.; Cho, C.H.; Sung, H.K. Design Parameter Optimization of a Silicon-Based Grating Waveguide for Performance Improvement in Biochemical Sensor Application. Sensors 2018, 18, 781. [Google Scholar] [CrossRef] [PubMed]
  27. Yan, S.; Zhang, M.; Zhao, X.; Zhang, Y.; Wang, J.; Jin, W. Refractive Index Sensor Based on a Metal–Insulator–Metal Waveguide Coupled with a Symmetric Structure. Sensors 2017, 17, 2879. [Google Scholar] [CrossRef] [PubMed]
  28. Gong, X.; Ngai, T.; Wu, C. A portable, stable and precise laser differential refractometer. Rev. Sci. Instrum. 2013, 84, 114103. [Google Scholar] [CrossRef] [Green Version]
  29. García-Valenzuela, A. Beam refraction and displacement in a differential refractometer with an absorbing sample. Opt. Lett. 2009, 34, 2192. [Google Scholar] [CrossRef]
  30. Zilio, S.C. A simple method to measure critical angles for high-sensitivity differential refractometry. Opt. Express 2012, 20, 1862. [Google Scholar] [CrossRef]
  31. McClimans, M.; LaPlante, C.; Bonner, D.; Bali, S. Real-time differential refractometry without interferometry at a sensitivity level of 10-6. Appl. Opt. 2006, 45, 6477. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, J.; Guo, W.; Xia, M.; Li, W.; Yang, K. In situ measurement of seawater salinity with an optical refractometer based on total internal reflection method. Opt. Express 2018, 26, 25510. [Google Scholar] [CrossRef] [PubMed]
  33. Sobral, H.; Peña-Gomar, M. Determination of the refractive index of glucose-ethanol-water mixtures using spectroscopic refractometry near the critical angle. Appl. Opt. 2015, 54, 8453. [Google Scholar] [CrossRef] [PubMed]
  34. Morales-Luna, G.; García-Valenzuela, A. Viability and fundamental limits of critical-angle refractometry of turbid colloids. Meas. Sci. Technol. 2017, 28, 125203. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, H.; Ye, J.; Yang, K.; Xia, M.; Guo, W.; Li, W. Real part of refractive index measurement approach for absorbing liquid. Appl. Opt. 2015, 54, 6046. [Google Scholar] [CrossRef] [PubMed]
  36. Lapsley, M.I.; Lin, S.C.S.; Mao, X.; Huang, T.J. An in-plane, variable optical attenuator using a fluid-based tunable reflective interface. Appl. Phys. Lett. 2009, 95, 083507. [Google Scholar] [CrossRef]
  37. Weber, E.; Vellekoop, M.J. Optofluidic micro-sensors for the determination of liquid concentrations. Lab Chip 2012, 12, 3754. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, L.; Zhang, Z.; Wang, Y.; Ye, M.; Fang, W.; Tong, L. Optofluidic refractive index sensor based on partial reflection. Photonic Sens. 2017, 7, 97–104. [Google Scholar] [CrossRef] [Green Version]
  39. Kedenburg, S.; Vieweg, M.; Gissibl, T.; Giessen, H. Linear refractive index and absorption measurements of nonlinear optical liquids in the visible and near-infrared spectral region. Opt. Mater. Express 2012, 2, 1588. [Google Scholar] [CrossRef]
  40. Daimon, M.; Masumura, A. High-accuracy measurements of the refractive index and its temperature coefficient of calcium fluoride in a wide wavelength range from 138 to 2326 nm. Appl. Opt. 2002, 41, 5275. [Google Scholar] [CrossRef]
  41. Polyanskiy, M.N. Refractive Index Database. Available online: https://refractiveindex.info (accessed on 2 February 2019).
  42. Dunn, M.H. Parametric Generation of Tunable Light from Continuous-Wave to Femtosecond Pulses. Science 1999, 286, 1513–1517. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Prism (left) and parallel plate (right) configurations, the latter shown N = 2 reflections at the sensing interface.
Figure 1. Prism (left) and parallel plate (right) configurations, the latter shown N = 2 reflections at the sensing interface.
Instruments 03 00036 g001
Figure 2. Transmittance T p (solid line) and relative intensity change Δ I / I for δ = δ n r = 1   μ RIU (dashed line) and δ = δ n i = 1   μ RIU (dotted line), versus incidence angle θ . Calculations assume n r / n 0 = 0.75 , n i / n 0 = 10 5 , N = 12 , α = 76.1 ° , n o = 2.16 , n a i r = 1 .
Figure 2. Transmittance T p (solid line) and relative intensity change Δ I / I for δ = δ n r = 1   μ RIU (dashed line) and δ = δ n i = 1   μ RIU (dotted line), versus incidence angle θ . Calculations assume n r / n 0 = 0.75 , n i / n 0 = 10 5 , N = 12 , α = 76.1 ° , n o = 2.16 , n a i r = 1 .
Instruments 03 00036 g002
Figure 3. Output signal Δ I / I c and angular acceptance bandwidth A B W , as a function of relevant variables. Open circles correspond to exact values derived from signal-versus-incidence-angle plots, by use of Equations (2) and (6); solid lines are estimates from the empirical Equations (7) and (9). Throughout, values of parameters other than x-axis variables are: n a i r = 1 , n r / n o = 0.75 , n i / n o = 10 5 , N / n o = 1 / 0.18 . Specifics on the index perturbation values δ , δ n r and δ n i are provided in each individual plot. A B W calculations assumed n o = 2.16 ; the angle α was properly varied to maintain a constant θ c = 15 ° in (g), (h,i), but a properly scalable θ c in (j). Respective values for Δ I / I c calculations are irrelevant, since they do not affect the results, but are reported for the sake of completeness: n o = 1.8 and α = 81 ° .
Figure 3. Output signal Δ I / I c and angular acceptance bandwidth A B W , as a function of relevant variables. Open circles correspond to exact values derived from signal-versus-incidence-angle plots, by use of Equations (2) and (6); solid lines are estimates from the empirical Equations (7) and (9). Throughout, values of parameters other than x-axis variables are: n a i r = 1 , n r / n o = 0.75 , n i / n o = 10 5 , N / n o = 1 / 0.18 . Specifics on the index perturbation values δ , δ n r and δ n i are provided in each individual plot. A B W calculations assumed n o = 2.16 ; the angle α was properly varied to maintain a constant θ c = 15 ° in (g), (h,i), but a properly scalable θ c in (j). Respective values for Δ I / I c calculations are irrelevant, since they do not affect the results, but are reported for the sake of completeness: n o = 1.8 and α = 81 ° .
Instruments 03 00036 g003
Figure 4. Sensitivity (left) and acceptance bandwidth (right), as a function of wavelength, for two indicative interfaces: C a F 2 / H 2 O (solid lines/open cicles) and S F 10 / H 2 O (dotted lines/black circles). Arrows indicate direction of increase in value of N ( = 1 , 4 , 16 ) and θ c (=15 deg, 75 deg), respectively. Open circles correspond to exact values derived from signal-versus-incidence-angle plots, by use of Equations (2) and (6); solid lines are estimates from the empirical Equations (8) and (9).
Figure 4. Sensitivity (left) and acceptance bandwidth (right), as a function of wavelength, for two indicative interfaces: C a F 2 / H 2 O (solid lines/open cicles) and S F 10 / H 2 O (dotted lines/black circles). Arrows indicate direction of increase in value of N ( = 1 , 4 , 16 ) and θ c (=15 deg, 75 deg), respectively. Open circles correspond to exact values derived from signal-versus-incidence-angle plots, by use of Equations (2) and (6); solid lines are estimates from the empirical Equations (8) and (9).
Instruments 03 00036 g004
Table 1. Results from the application of the optimization protocol to three hypothetical scenarios. Calculations assume δ 1 = δ 2 = 1 , M 2 = 1.2 and σ d e t = 0 . Units: T in °C; λ in μ m; A B W in degrees; w o in mm; D L in μ RIU. Other quantities are dimensionless. Sensor’s linear range can be deduced from available data by use of Equation (16). Minimum useful signal (i.e., relative intensity change) at the detector amounts to S · D L 4 % , 2 % and 0.4 % , for Cases 1, 2 and 3, respectively.
Table 1. Results from the application of the optimization protocol to three hypothetical scenarios. Calculations assume δ 1 = δ 2 = 1 , M 2 = 1.2 and σ d e t = 0 . Units: T in °C; λ in μ m; A B W in degrees; w o in mm; D L in μ RIU. Other quantities are dimensionless. Sensor’s linear range can be deduced from available data by use of Equation (16). Minimum useful signal (i.e., relative intensity change) at the detector amounts to S · D L 4 % , 2 % and 0.4 % , for Cases 1, 2 and 3, respectively.
ParametersCase 1Case 2Case 3
Δ T 1 0.1 0.01
Δ λ 10 3 10 4 10 5
σ l a s e r 1 % 0.5 % 0.1 %
n r 1.5 1.5 1.5
n i 3 × 10 3 1.3 × 10 4 1.0 × 10 5
d n r / d T 10 4 5 × 10 5 10 5
d n r / d λ 0.1 0.05 0.01
n o 2 1.8 1.6
N124
θ c 154575
A B W 1.82 0.12 0.03
A B W / θ d 41.5 · w o / λ 2.7 · w o / λ 0.8 · w o / λ
S = S o p t 717107100
D L 60030 0.6

Share and Cite

MDPI and ACS Style

Koutsoumpos, S.; Giannios, P.; Triantis, D.; Moutzouris, K. Critical-Angle Differential Refractometry of Lossy Media: A Theoretical Study and Practical Design Issues. Instruments 2019, 3, 36. https://doi.org/10.3390/instruments3030036

AMA Style

Koutsoumpos S, Giannios P, Triantis D, Moutzouris K. Critical-Angle Differential Refractometry of Lossy Media: A Theoretical Study and Practical Design Issues. Instruments. 2019; 3(3):36. https://doi.org/10.3390/instruments3030036

Chicago/Turabian Style

Koutsoumpos, Spyridon, Panagiotis Giannios, Dimos Triantis, and Konstantinos Moutzouris. 2019. "Critical-Angle Differential Refractometry of Lossy Media: A Theoretical Study and Practical Design Issues" Instruments 3, no. 3: 36. https://doi.org/10.3390/instruments3030036

Article Metrics

Back to TopTop