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

Concurrent polarization IR analysis to determine the 3D angles and the order parameter for molecular orientation imaging

Open Access Open Access

Abstract

A non-tomographic analysis method is proposed to determine the 3D angles and the order parameter of molecular orientation using polarization-dependent infrared (IR) spectroscopy. Conventional polarization-based imaging approaches provide only 2D-projected orientational information of single vibrational modes. The newly proposed method concurrently analyses polarization angle-dependent absorptance of two non-parallel transition dipole moments. The relative phase angle and the maximum-to-minimum ratios observed from the two polarization profiles are used to calculate the 3D angles of the mean molecular orientation and the order parameter of the orientational distribution. Usage of those relative observables as intermediate input parameters makes the analysis results robust against variations in concentration, thickness, absorption peak, and absorption cross-section, which can occur in typical imaging conditions. This analysis is based on a single-step, non-iterative calculation that does not require any analytical model function of an orientational distribution function. This concurrent polarization analysis method is demonstrated using two simulation data examples, followed by associated error propagation analysis and discussion on the effect of absorption strength. Application of this robust spectral analysis method to polarization IR microscopy will provide a full molecular orientation image without tilting that tomographies require.

1. Introduction

Molecular anisotropy plays a critical role in nature, affecting not only the directionality of a bulk property but also creating new properties different from isotropic materials. Accurate measurement and proper characterization of the spatial heterogeneity and hierarchy in molecular orientation is important in the fundamental understanding of the effects of structured molecular anisotropy on the biological, chemical, and mechanical properties of macromolecular materials [1–3]. Various spectroscopic and optical imaging techniques for orientation measurement are mostly based on polarization-controlled spectroscopy, such as IR absorption [3–5], fluorescence [6], confocal Raman [7], coherent Raman [8,9], and second harmonic generation [10]. Despite the diffraction-limited spatial resolution, these methods can only measure the molecular orientation projected onto the polarization rotating plane and the in-plane order parameters. 2D-projected orientation imaging can provide quantitative information for samples where the symmetry axis is pre-defined onto the polarization rotating plane, such as fibers [11] and stretched films [12]. However, for samples whose orientation can be out-of-plane, only qualitative information can be obtained by 2D-projected imaging methods. Recently, different imaging approaches have been developed to determine the out-of-plane component of molecular orientation by analyzing z-stacked images [13–16], interference patterns or intensity distributions of single molecule emissions [17–19], but those methods are applicable only to isolated chromophores or defects; thus, not appropriate for imaging of continuous materials. Alternatively, 3D orientation can be accurately determined by tomographical methods by tilting the sample while characterizing [20–22]. Despite their successful demonstrations for 3D orientation measurement of molecular orientation in polymer thin films, those tomographic approaches are very challenging to apply to high-resolution optical imaging due to spatial restriction of high magnification objective lenses. Therefore, these optical tomographic approaches have been demonstrated mostly for homogeneous thin samples.

Molecular orientation can be described either in the laboratory coordinate or in the molecular coordinate. For mathematical simplicity, the mean orientation direction is described with the 3D angles in the laboratory coordinate while the local orientational broadening is described in the molecular coordinate with respect to the mean orientation direction.

Molecular orientation is mathematically expressed by an orientational distribution function (ODF). For a molecular system with cylindrical symmetry, the ODF can be expressed by a series of Legendre polynomials with respect to the symmetry axis. Depending on the optical interaction mechanism, each measurement can determine a different number of coefficients in the Legendre polynomials [1,23]. For example, wide angle X-ray scattering (WAXS) is known to determine all coefficients. Nuclear magnetic resonance (NMR) can measure up to the eighth-order coefficients. Raman and fluorescence can provide the second- and fourth-order coefficients, and infrared (IR) absorption and second harmonic generation can measure the second-order coefficient [2,23]. Among the coefficients, in particular, the second-order coefficient, also known as the order parameter, is widely used to represent the broadening of an ODF. In this paper, the order parameter is used as a metric of orientational broadening of the ODF in the molecular coordinate.

In the previous paper [24], a new spectral analysis method was proposed to determine the 3D angles of molecular orientation, analyzing coherent anti-Stokes Raman scattering (CARS) susceptibilities of two non-parallel Raman modes observed by a single polarization scan. In the paper, the spectral analysis method determines the 3D orientation angles of molecular systems with different symmetries and orientational broadening. However, the previous approach requires an analytical model function for the ODF to determine the mean orientation angles and the order parameter of the ODF. Its calculation is based on iterative computation, which potentially costs expensive computation time.

In this paper, a different approach based on polarization IR is proposed to determine the 3D angles and the orientational order parameter without assuming any model function for an ODF. This concurrent polarization IR analysis method can also calculate those output values by a single step without iterative computation. Similar to the approach developed for polarization CARS [24], this new IR analysis method takes advantage of dimensionless parameters observed from two polarization IR absorption profiles. The concurrent polarization IR analysis method can be easily implemented to conventional IR microscopy by a simple addition of polarization rotating optics in the beam path.

2. Theory

2.1. Polarization angle-dependent absorptance by aligned transition dipole moments

In literature, polarization angle dependence of light absorption by aligned molecules has been described with either absorbance A ≡ − logT [25–28], or absorptance α ≡ 1 − T [29–31], where T is the transmittance. However, those expressions were used without thorough derivations or clarifications of underlying assumptions. Here, polarization angle-dependent light absorption is described for aligned molecules using classical electromagnetic wave theories. Figure 1 shows how linearly polarized light intensity is reduced by a collection of transition dipole moments that are aligned parallel to each other. In this scheme, the incident electric field is divided into the parallel and perpendicular electric fields with respect to the plane defined by the light propagation direction and the transition dipole moment direction. While the parallel electric field is reduced through the sample, the perpendicular electric field is unaffected. As derived in Fig. 1, the transmitted light intensity, I(n), is expressed with the incident light intensity, I(0), the polarization angle, η, and the transition dipole tilting angle, θ, as

I(d)=I(0){1[1exp(2εcdsin2θ)]cos2(ηψ)}
where ψ is the azimuthal angle of the transition dipole moments on the xy plane. Then, the polarization angle-dependent transmittance can be expressed as
T(η)I(d)/I(0)=1[1exp(2εcdsin2θ)]cos2(ηψ)
Using Eq. (2) and the relation of A = − logT, the absorbance can also be expressed as a function of η, but in a complex form as
A(η)=log{1[1exp(2εcdsin2θ)]cos2(ηψ)}
which is far from the expression of cos2 η for absorbance, which is frequently used in literatures. In contrast, Eq. (2) can be rewritten into a simpler form of expression using absorptance, α = 1 − T as
α(η)=[1exp(2εcdsin2θ)]cos2(ηψ)
which shows that α(η) ∝ cos2(ηψ) and that the phase of α(η) corresponds to the azimuthal angle ψ. However, the tilting angle, θ, is still blended with the non-orientational quantities in the exponent, which makes it difficult to extract the orientation angles.

 figure: Fig. 1

Fig. 1 Polarization angle dependence of transmitted light intensity through a collection of aligned transition dipole moments. The incident light, I(0), propagating along the z axis, is linearly polarized with the polarization angle rotated by η from the x axis in the xy plane. All transition dipole moments are aligned parallel in the xz plane with the tilting angle of θ from the z axis. The magnitude of the transition dipole moment, equivalent to the absorption cross section, is represented by ε. The terms d and c denote the sample thickness and the concentration, respectively. As the light passes through the sample, only Ex is reduced by absorption but Ey is unaffected.

Download Full Size | PDF

A Taylor series is a commonly used representation that transforms a nonlinear function into a series of polynomial terms. If the exp() term in Eq. (4) is represented by a Taylor series, Eq. (4) can be rewritten as

α(η)={1[1(2εcdsin2θ)+(2εcdsin2θ)2/2]}cos2(ηψ)=[(2εcdsin2θ)(2εcdsin2θ)2/2+]cos2(ηψ).
When (2εcd sin2θ) ≪ 1, or when light absorption is weak, the polarization angle-dependent absorptance, α(η), can be simplified as
α(η)(2εcd)sin2θcos2(ηψ)α°sin2θcos2(ηψ)
where α° ≡ (2εcd) contains all non-orientational quantities. In the simplified form of α(η) in Eq. (6), α° corresponds to the maximum absorptance that can be measured when all transition dipole moments are parallel to the incident light polarization direction. Later in the last discussion section, more quantitative description is provided about the validness of the weak-absorption approximation in terms of the analysis results.

It is noted that in Eq. (6), the non-orientational term [α° = 2εcd] is separated from the orientational term [sin2 θ cos2(ηψ)]. Interestingly, the orientational term can be alternatively expressed as the inner product of the unit vector of the light polarization, êE, and the directional unit vector of the transition dipole moment, êμ, as

α(η)=α°[e^E(η)e^μ(ψ,θ)]2
where êE(η) = (cos η, sin η, 0) and êμ(ψ, θ) = (cos ψ sin θ, sin ψ sin θ, cos θ). This expression of absorptance as the inner product is useful for an analytical description of absorptance using an ODF in the next section.

2.2. Polarization profiles of orientationally broadened transition dipole moments

Unlike conventional polarization analysis methods, this newly proposed analysis method uses polarization-dependent absorption profiles of two non-parallel transition dipole moments. Figure 2 shows a schematic presentation of broadening and rotation of a molecule represented by two orthogonal transition dipole moments. The primary transition dipole moment μ1 represents the main molecular axis with a high symmetricity, e.g., the main chain axis of a polymer. The secondary transition dipole moment μ2 represents a different vibrational mode that is orthogonal to μ1. If μ1 and μ2 are assumed to be broadened by the same degree, their local ODF with respect to the mean direction of each transition dipole moment will be represented by the same local ODF, ρ(β). Rotation of a molecule from the molecular coordinate into the laboratory coordinate is represented by the Euler rotation matrix as

R(ψ,θ,ϕ)=(cosψsinψ0sinψcosψ0001)(cosθ0sinθ010sinθ0cosθ)(cosϕsinϕ0sinϕcosϕ0001)
where ψ, θ, and ϕ are the azimuthal, axial, and rotational angles. Then, from Eq. (7), absorptance α1(η) for μ1 can be expressed as
α1(η)=α1°02π0π[e^E(η)e^μ1]2ρ(β)sinβdβdγ=α1°02π0π[e^E(η)R(ψ,θ,ϕ)R(γ,β,0)e^z]2ρ(β)sinβdβdγ
where êz = (0, 0, 1) is the unit vector of original μ1 direction in the molecular coordinate before broadening and rotation; and β and γ are the axial and rotational angles in the molecular coordinate, respectively.

 figure: Fig. 2

Fig. 2 Schematic presentation of broadening and rotation of molecular orientation. The orientational broadening is represented by ρ(β), where β denotes the axial angle of an individual μ from the mean orientation direction, 〈μ〉. The molecular rotation is represented by the rotation matrix R(ψ, θ, ϕ) from the molecular coordinate to the laboratory coordinate. The incident light propagates in the direction of the z axis, and the light polarization plane is on the xy plane. η is the polarization angle with respect to the x axis.

Download Full Size | PDF

If the local ODF is cylindrically symmetrical, ρ(β) is independent of γ. Then, integration of Eq. (9) with respect to γ becomes:

α1(η)α1°=12[1sin2θcos2(ηψ)]+12[3sin2θcos2(ηψ)1]0πcos2βρ(β)sinβdβ

Derivation of the absorptance of the secondary transition dipole moment, α2(η), is similar to that of α1(η), but with an additional step. First, the secondary transition dipole moments is set along the z axis and broadens by ρ(β) with respect to the z axis. The broadened α2(η) is rotated by π2 around the y axis toward the x axis, followed by rotation by R(ψ, θ, ϕ) into the laboratory coordinate. The analytical form of α2(η) becomes:

α2(η)=α2°02π0π[e^E(η)R(ψ,θ,ϕ)Ry(π2)R(γ,β,0)e^z]2ρ(β)sinβdβdγ
where Ry(π2) denotes the rotation matrix around the y axis by π2. After the integration with respect to γ, Eq. (11) becomes:
α2(η)α2°=12{[cosθcosϕcos(ηψ)+sinϕsin(ηψ)]21}+12{3[cosθcosϕcos(ηψ)+sinϕsin(ηψ)]21}×0πcos2βρ(β)sinβdβ

For visualization of the polarization-dependent absorptance profiles of α1(η) and α2(η), an analytical function is used for ρ(β) in Eqs. (10) and (12). It should be noted that this analytical model function is used simply for data simulation, not used for data analysis. A von Mises-Fisher distribution function, ρvMF(β), shown in Fig. 3a, is widely used to represent a cylindrical ODF [24]. A von Mises-Fisher distribution function monotonically decreases from the maximum at β = 0, and its orientation broadening is represented by a parameter, called the concentration parameter, κ. When κ increases, ρvMF(β) becomes narrower. For example, when κ approaches to ∞, the ODF has no broadening. Figure 3a shows two examples of ρvMF(β) plotted for two κ values. Using the von Mises-Fisher distribution function for ρ(β), polarization-dependent absorptance profiles α1(η) and α2(η) are calculated from Eqs. (10) and (12) for two different Euler rotation angles, as shown in Figs. 3b and 3c. For comparison, they show α1(η) and α2(η) with an infinitely narrow ODF, represented as ρvMF(β) with. In the presence of broadening, the minimum absorptance becomes non-zero. The maximum values of α1(η) and α2(η) can either increase or decrease, depending on (ψ, θ, ϕ) and κ. These simulated polarization profiles are used for demonstration of 3D angles calculation and error analysis later in this paper.

 figure: Fig. 3

Fig. 3 (a) Plots of the von Mises-Fisher function ρvMF(β) for two different κ values. (b) and (c) Plots of absorptance calculated from Eqs. (10) and (12) for the primary (red) and the secondary (blue) transition dipole moments using two different sets of orientation angles of the mean orientation directions. The 3D angles used for calculation are indicated above each figure. For comparison, absorptance profiles from unbroadened ODF (κ = ∞) are plotted as the solid lines. The dashed lines indicate absorptance profiles calculated from broadened ρvMF(β) with κ = 10 (b) and κ = 50 (c). For both (b) and (c), α1°=0.9 and α2°=0.9.

Download Full Size | PDF

2.3. Legendre polynomials for an orientation distribution function

For molecular orientation with cylindrical symmetry, the ODF can be expressed with a series of Legendre polynomials, Pn(cos β), and the coefficients, 〈Pn〉 as

ρ(β)=n=0(n+12)PnPn(cosβ).
Because of the symmetry with respect to the mirror plane perpendicular to the principal axis, all the odd-order terms in Eq. (13) are equal to zero. In addition, polarization-dependent IR absorptance signals are associated only with P0(cos β) and P2(cos β) terms because of the square dependence of the absorptance on the product of the electric field and the transition dipole moment [2,23]. The first two even-order Legendre polynomials, associated with polarization IR, are:
P0(cosβ)=1
P2(cosβ)=12(3cos2β1).
While the zero-order coefficient, 〈P0〉, is always equal to one, the second-order coefficient, 〈P2〉, varies from 13 to 1. 〈P2〉, also called as the order parameter, is used as a metric of the breadth of an ODF. For an infinitely narrow distribution parallel to the symmetry axis, 〈P2〉 is equal to one. For an isotropic distribution, 〈P2〉 is equal to zero. For a distribution where all orientations are perpendicular to the symmetry axis, 〈P2〉 becomes equal to 13.

Now that an ODF is expressed as a generalized form of Legendre polynomials, the absorptance of the primary and secondary transition dipole moments in Eqs. (10) and (12), respectively, can be expressed by the Legendre polynomial coefficients. The integral term common in Eqs. (10) and (12) can be rewritten as

0πcos2βρ(β)sinβdβ=2P2+13.
By inserting Eq. (16) into Eq. (10), the absorptance of the primary transition dipole moment is expressed as
α1(η)α1°=P2sin2θcos2(ηψ)+13(1P2).
Similarly, by inserting Eq. (16) into Eq. (12), the absorptance of the secondary transition dipole moment is calculated as
α2(η)α2°=P2[cosθcosϕcos(ηψ)+sinϕsin(ηψ)]2+13(1P2).

In a conventional non-tomographic approach to polarization scanning spectroscopy, the apparent order parameter, 〈P′2〉, is often determined by the maximum and minimum values of a polarization profile [5,12,32]. Using Eq. (17), the maximum and minimum absorptances, αmax and αmin, respectively, are calculated and used to relate the apparent order parameter, 〈P′2〉, with the original order parameter,

P2=αmaxαminαmax+2αmin=P2sin2θ1P2cos2θ
Equation (19) shows that 〈P′2〉 becomes equal to 〈P2〉 only when θ = π/2. This is why the apparent 〈P′2〉 determined by the maximum and minimum values of a polarization-dependent measurement can be meaningful only when the mean orientation direction is already known to be parallel to the polarization plane; e.g., stretched films [12] and fibers [5,11]. However, if the mean direction of the molecular axis is unknown from the polarization plane, the apparent 〈P′2〉 cannot be used as the order parameter of the original ODF. This clearly shows that measurement of the order parameter is not separable from a measurement of the 3D angle of the mean molecular orientation.

2.4. Calculation of 3D angles (ψ, θ, ϕ) and order parameter 〈P2

This section describes how to determine the 3D orientation angles (ψ, θ, and ϕ) and the order parameter 〈P2〉 using α1(η) and α2(η). First, the azimuthal angle ψ of the molecular main axis can be directly measured from the phase of the α1(η) profile, as shown in Eq. (17),

ψ=η1,max(0ψπ)
where η1,max is the polarization angle at α1,max. It is noted that because of inversion symmetry of the electric field of linearly polarized light, this linear polarization-based measurement can determine the azimuthal angle ψ only in the range of from 0 to π and cannot determine the absolute phase of the orientation. The other two angles, θ and ϕ, cannot be directly measured from a single polarization profile, as discussed above. To calculate the other two angles, three intermediate parameters are defined using the phase angles and the maximum and minimum values of both polarization profiles.

The phase difference is defined as Δηη2,maxη1,max, and its relation with θ and ϕ can be analytically expressed using the solution of dα2(η)dη=0 from Eq. (18), as

cosϕcosθsinΔη=sinϕcosΔη
Δη=tan1(tanϕsecθ)

The maximum and minimum values of α1(η) and α2(η) are expressed with θ, ϕ, and 〈P2〉 from Eqs. (17) and (18) as

α1,minα1°=13(1P2)
α1,maxα1°=P2sin2θ+13(1P2)
α2,minα2°=13(1P2)
α2,maxα2°=P2(cos2ϕcos2θ+sin2ϕ)+13(1P2)
Then, a modified form of the ratio of the minimum and maximum absorptance values of α1(η) is defined for the primary mode as M
M(α1,maxα1,min1)=3P21P2×sin2θ
Similarly, for the secondary mode, a modified form of the ratio of the minimum and maximum absorptance values N is defined as
N(α2,maxα2,min1)=3P21P2(cos2ϕcos2θ+sin2ϕ)

Using these three intermediate parameters (Δη, M, and N), the three remaining unknown variables, θ, ϕ, and 〈P2〉, can be obtained from the three relations of Eqs. (22), (27), and (28). For simplifying the derivation, a temporary variable, P, is defined on behalf of 〈P2〉 as

P3P21P2
Then, Eq. (27) can be rewritten as
sin2θ=M/P
By inserting Eq. (30) into Eq. (28),
cos2ϕ=(PN)/M
From Eqs. (21) and (28),
(N/P)cos2Δη=cos2ϕcos2θ=cos2ϕ(1sin2θ)
Insertion of Eqs. (30) and (31) into Eq. (32) leads to a quadratic equation with respect to P as
(PM)(PN)=MNcos2Δη
The solution of the quadratic equation is:
P=12[(M+N)±(M+N)24MN(1cos2Δη)]=12[(M+N)±(MN)2+4MNcos2Δη]
When this solution is inserted into Eq. (31),
cos2ϕ=12M[(MN)±(MN)2+4MNcos2Δη]0
Because both M and N are positive, the condition cos2ϕ ≥ 0 can be met only with the positive sign in front of the square root term in Eq. (35) as
cos2ϕ=12M[(MN)+(MN)2+4MNcos2Δη]
By inserting Eq. (36) back into Eq. (31), P can be expressed as a single-signed value as
P=12[(M+N)+(MN)2+4MNcos2Δη]

Now that P can be determined with the three observables, Δη, M, and N, Eq. (37) can be inserted into Eqs. (29), (30), and (31), providing the values of 〈P2〉, θ, and ϕ, respectively.

P2=PP+3
θ=sin1M/P(0θπ)
ϕ=cos1[±(PN)/M](πϕπ)
The sign in the brace of Eq. (40) is determined from the sign of tan Δη, from Eq. (21). It is important to know that a given set of observables (Δη, M, and N) will generate two sets of solutions for θ and ϕ from Eqs. (21) and (39). If the first solution pair is set as 0 ≤ θ1π/2 and 0 ≤ ϕ1π, the second pair is its mirror image to the xy plane, as θ2 = πθ1 and ϕ2 = −ϕ1. The existence of two solution pairs for θ and ϕ is because the transition dipole moments with a reflection symmetry to each other with respect to the polarization plane are indistinguishable in their polarization profiles.

Figure 4 summarizes how the six observables from the polarization profiles of two orthogonal IR absorption modes are used to simultaneously determine all 3D angles, ψ, θ, and ϕ, and the order parameter, 〈P2〉. It should be noted that the intermediate parameters, M, N, and Δη, are dimensionless and are not affected by variation in sample conditions, such as concentration, thickness, absorption frequency, and absorption cross section. It should also be noted that the 〈P2〉 value determined by this method is for the original molecular-coordinate ODF with respect to the mean molecular orientation direction, which is not affected by the out-of-plane tilting.

 figure: Fig. 4

Fig. 4 A flow chart of the concurrent polarization analysis method for determination of the 3D angles (ψ, θ, and ϕ) of the mean molecular orientation and the order parameter 〈P2〉 of the local orientation distribution using polarization profiles of two orthogonal IR modes. The intermediate parameters (Δη, M, and N) are calculated from the six input observables of the phase angles and the maximum and minimum absorptances of the two polarization profiles.

Download Full Size | PDF

2.5. Error analysis

The earlier sections show that the analytical solutions of the 3D angles and the order parameter of molecular orientation can be obtained from a simple calculation with polarization dependent absorptances observed from two orthogonal IR modes. However, experimental data contains uncertainty or unwanted error in the measurement due to imperfect polarization optics, detector noise, and other system fluctuations. In this section, we examine how the errors in the input observables propagate to the intermediate parameters and to the outputs of the 3D angles and the order parameter. The uncertainty of each variable is represented by the standard deviation. The uncertainty of an output value is calculated from the uncertainties of the input values using the following error propagation relation assuming that there is no correlation between any input observables,

δf(x1,,xn)=(fx1)2(δx1)2++(fxn)2(δxn)2.

As expected from the complex relations among the input observables and the output variables, their error propagations are non-linear and complex. Instead of listing all error propagation equations of each output variable, the uncertainties in the output values are examined using simulation data generated with known input variables. The two examples of polarization profiles previously presented in Figs. 3b and 3c are used for this uncertainty analysis.

In actual experiments, measured transmittance may include non-negligible contributions from reflection from interfaces or scattering, resulting in an overestimation of absorptance [33]. Because those factors can be both wavelength- and polarization-dependent, one may need to consider their effects into an analysis of polarization-IR absorptance measurements although those contributions are not discussed in this paper.

Table 1 shows how the input observables are converted to the intermediate parameters and then to the output values. The results confirm that the output results of this calculation correspond to the original input values used to generate the tested simulated data. It is also noted that the 3D angles of the mirror image of the original molecular orientation are also calculated as a complementary solution. In Table 1, the uncertainties (the standard deviations) of the intermediate parameters and the output values are computed based on the error propagation relation of Eq. (41). The uncertainty in ψ is identical to the uncertainty in the input η1,max from Eq. (20). Calculation of the uncertainties of ϕ and θ is more complicated but their output uncertainties are in the same order of magnitude with the input uncertainties. In these two specific examples, the output uncertainty in 〈P2〉 is found to be 0.02 or lower, which can be considered as very small, compared to the whole range of 13P21. It should be noted that the calculated output uncertainties are for these two specific simulation data, and the resulting uncertainty will vary non-linearly with the input observables as well as their uncertainties.

Tables Icon

Table 1. Demonstration of determination of the 3D angles and 〈P2〉 using observables and intermediate parameters from the simulated polarization absorptance profiles shown in Fig. 3. The input uncertainties for phase angle η and absorptance α are assumed to be 2° and 5%, respectively, for Fig. 3b, and 5° and 10% for Fig. 3c. The parentheses for θ and ϕ indicate the second solution for the 3D angles of the mirror image molecular orientation with respect to the polarization plane.

2.6. Validity of the weak absorption approximation

Earlier in Eq. (4), polarization angle-dependent absorptances, α(η), of all parallel transition dipole moments is expressed as a cos2(ηψ) multiplied by an amplitude containing an exponential function. Then, the original form of α(η) is simplified as a product of a non-orientational term and an orientational term using the weak absorption approximation. The simplified form of α(η) is used to derive the following ODF-broadened absorptance into Eqs. (10) and (12) for the primary and the secondary modes, respectively. Here, the weak absorption approximation is examined in terms of the analysis outputs of 3D orientation angles and order parameters. For simplicity, the ODF-broadened absorptances calculated from the simple and the original forms of α(η) are denoted as αsimplifiedODF(η) and αoriginalODF(η), respectively. Unlike αsimplifiedODF(η), αoriginalODF(η) cannot be expressed as a simple analytical equation without integrals. Therefore, a numerical calculation is performed using a von Mises-Fisher function as a model ODF for αoriginalODF(η).

Figure 5 shows numerically calculated αoriginalODF(η) for three different α° values, which are compared with the corresponding αsimplifiedODF(η). The shapes of the αoriginalODF(η) curves correspond to an offset cosine function, which is similar to the shape of αsimplifiedODF(η). Also, the phases are identical between αoriginalODF(η) and αsimplifiedODF(η). The only difference between the two curves is the amplitude and the offset. As α° increases, αsimplifiedODF(η) increases greater than αoriginalODF(η), exhibiting the inadequacy of the weak absorption approximation for a large α° case. However, the concurrent polarization analysis introduced in this paper is not performed based on the absolute absorptance values but on the phase angles and the maximum-to-minimum ratios, M and N. Interestingly, M values calculated from αoriginalODF(η) and αsimplifiedODF(η) are very close to each other even when α° is close to one. The closeness in M between αoriginalODF(η) and αsimplifiedODF(η) indicates that the formulation derived using the weak absorption approximation can still be useful for the concurrent polarization analysis. In addition, both αoriginalODF(η) and αsimplifiedODF(η) are not affected by α°, which suggests that the 3D angles and the order parameter determined by the analysis are not strongly influenced by variation in sample thickness, concentration, and absorption cross section, even when α° is large. In Eq. (27), when α° increases from 0.01 to 0.99 (corresponding to optical densities from 0.004 to 2), the relative variations in M values are as low as 1.6% and 4% for θ = 60° and 30°, respectively. This difference is comparable to or smaller than typical experimental uncertainty values in M. This unexpected insensitivity of intermediate parameters to sample variations even at a strong absorption condition demonstrates that the concurrent polarization analysis method based on dimensionless parameters can be effective for a wide range of absorptance.

 figure: Fig. 5

Fig. 5 (a), (b), (c) Comparison of polarization angle-dependent absorptance calculated for ODF-broadened orientation based on the simplified and the original forms of α(η). The calculation is performed for the primary mode, and the axial angle θ = 60°. The von Mises-Fisher function with κ = 10 is used as the model ODF for the numerical calculation. (d), (e) The maximum-to-minimum ratio, represented by M in Eq. (27) is plotted as a function of α° for different θ.

Download Full Size | PDF

3. Conclusion

A new non-tomographic approach based on polarization IR spectroscopy has been described to measure the 3D angles of the mean molecular orientation and the order parameter of the local molecular orientation. Six input observables from polarization profiles of two non-parallel transition dipole moments are converted into three intermediate, dimensionless parameters. Then, a single-step, non-iterative calculation using these parameters determines the 3D Euler angles (ψ, θ, and ϕ) and the order parameter, 〈P2〉, without requiring any analytical model function for the orientational distribution function. Usage of the relative intermediate parameters makes this concurrent polarization IR analysis method insensitive to variations in the sample concentration, thickness, and absolute absorption cross section. When applied to polarization IR microscopy, this concurrent polarization analysis method will provide a full orientational image map with a diffraction limited spatial resolution without tilting that tomographies require. The formulation in the paper is targeted for polarization IR imaging, but the same principle of this polarization analysis method can be applied to Raman imaging, which would expand the application systems of 3D orientation imaging.

Acknowledgments

The author thanks Isao Noda for comments on polarization angle dependence of absorbance and absorptance. Official contribution of the National Institute of Standards and Technology; not subject to copyright in the United States.

References and links

1. D. I. Bower, “Orientation distribution functions for uniaxially oriented polymers,” J. Polym. Sci. Polym. Phys. Ed. 19, 93–107 (1981). [CrossRef]  

2. M. Tanaka and R. J. Young, “Review Polarised Raman spectroscopy for the study of molecular orientation distributions in polymers,” J. Mater. Sci. 41, 963–991 (2006). [CrossRef]  

3. Y. Hikima, J. Morikawa, and T. Hashimoto, “Wavenumber dependence of FT-IR image of molecular orientation in banded spherulites of poly(3-hydroxybutyrate) and poly(l-lactic acid),” Macromolecules. 46, 1582–1590 (2013). [CrossRef]  

4. T. P. Wrobel, P. Mukherjee, and R. Bhargava, “Rapid visualization of macromolecular orientation by discrete frequency mid-infrared spectroscopic imaging,” Analyst. 142, 75–79 (2016). [CrossRef]   [PubMed]  

5. M. Ryu, A. Balcytis, X. Wang, J. Vongsvivut, Y. Hikima, J. Li, M. J. Tobin, S. Juodkazis, and J. Morikawa, “Orientational Mapping Augmented Sub-Wavelength Hyper-Spectral Imaging of Silk,” Sci. Reports 7, 7419 (2017). [CrossRef]  

6. A. T. Ngo, Z. J. Jakubek, Z. Lu, B. Joós, C. E. Morris, and L. J. Johnston, “Membrane order parameters for interdigitated lipid bilayers measured via polarized total-internal-reflection fluorescence microscopy,” Biochimica et Biophys. Acta - Biomembr. 1838, 2861–2869 (2014). [CrossRef]  

7. F. L. Labarthet, J.-L. Bruneel, T. Buffeteau, C. Sourisseau, M. R. Huber, S. J. Zilker, and T. Bieringer, “Photoinduced orientations of azobenzene chromophores in two distinct holographic diffraction gratings as studied by polarized Raman confocal microspectrometry,” Phys. Chem. Chem. Phys. 2, 5154–5167 (2000). [CrossRef]  

8. S. H. Parekh and K. F. Domke, “Watching orientational ordering at the nanoscale with coherent anti-stokes raman microscopy,” Chem. - A Eur. J. 19, 11822–11830 (2013). [CrossRef]  

9. C. Cleff, A. Gasecka, P. Ferrand, H. Rigneault, S. Brasselet, and J. Duboisset, “Direct imaging of molecular symmetry by coherent anti-Stokes Raman scattering,” Nat. Commun. 7, 11562 (2015). [CrossRef]  

10. P. Stoller, K. M. Reiser, P. M. Celliers, and A. M. Rubenchik, “Polarization-Modulated Second Harmonic Generation in Collagen,” Biophys. J. 82, 3330–3342 (2002). [CrossRef]   [PubMed]  

11. M. Richard-lacroix and C. Pellerin, “Accurate New Method for Molecular Orientation Quantification Using Polarized Raman Spectroscopy,” Macromolecules. 46, 5561–5569 (2013). [CrossRef]  

12. Y. Hikima, J. Morikawa, and T. Hashimoto, “FT-IR image processing algorithms for in-plane orientation function and azimuth angle of uniaxially drawn polyethylene composite film,” Macromolecules. 44, 3950–3957 (2011). [CrossRef]  

13. E. Toprak, J. Enderlein, S. Syed, S. A. McKinney, R. G. Petschek, T. Ha, Y. E. Goldman, and P. R. Selvin, “Defocused orientation and position imaging (DOPI) of myosin V,” Proc. Natl. Acad. Sci. 103, 6495–6499 (2006). [CrossRef]   [PubMed]  

14. N. Sandeau, L. Le Xuan, D. Chauvat, C. Zhou, J.-F. Roch, and S. Brasselet, “Defocused imaging of second harmonic generation from a single nanocrystal,” Opt. Express 15, 16051–16060 (2007). [CrossRef]   [PubMed]  

15. E. A. Büyüktanir, K. Zhang, A. Gericke, and J. L. West, “Raman Imaging of Nematic and Smectic Liquid Crystals,” Mol. Cryst. Liq. Cryst. 487, 39–51 (2008). [CrossRef]  

16. A. V. Kachynski, A. N. Kuzmin, P. N. Prasad, and I. I. Smalyukh, “Coherent anti-Stokes Raman scattering polarized microscopy of three-dimensional director structures in liquid crystals,” Appl. Phys. Lett. 91, 151903 (2007). [CrossRef]  

17. S. A. Empedocles, R. Neuhauser, and M. G. Bawendi, “Three-dimensional orientation measurements of symmetric single chromophores using polarization microscopy,” Nature. 399, 126–130 (1999). [CrossRef]  

18. A. P. Bartko and R. M. Dickson, “Imaging Three-Dimensional Single Molecule Orientations,” J. Phys. Chem. B 103, 11237–11241 (1999). [CrossRef]  

19. R. Börner, N. Ehrlich, J. Hohlbein, and C. G. Hübner, “Single Molecule 3D Orientation in Time and Space: A 6D Dynamic Study on Fluorescently Labeled Lipid Membranes,” J. Fluoresc. 26, 963–975 (2016). [CrossRef]   [PubMed]  

20. T. Elzein, M. Nasser-Eddine, C. Delaite, S. Bistac, and P. Dumas, “FTIR study of polycaprolactone chain organization at interfaces,” J. Colloid Interface Sci. 273, 381–387 (2004). [CrossRef]   [PubMed]  

21. A. M. Anton, R. Steyrleuthner, W. Kossack, D. Neher, and F. Kremer, “Infrared Transition Moment Orientational Analysis on the Structural Organization of the Distinct Molecular Subunits in Thin Layers of a High Mobility n-Type Copolymer,” J. Am. Chem. Soc. 137, 6034–6043 (2015). [CrossRef]   [PubMed]  

22. W. Kossack, M. Schulz, T. Thurn-Albrecht, J. Reinmuth, V. Skokow, and F. Kremer, “Temperature-dependent IR-transition moment orientational analysis applied to thin supported films of poly-∊-caprolactone,” Soft Matter 13, 9211–9219 (2017). [CrossRef]   [PubMed]  

23. C.-P. Lafrance, A. Nabet, R. E. Prud’homme, and M. Pezolet, “On the relationship between the order parameter (P2(cos θ)) and the shape of orientation distributions,” Can. J. Chem. 73, 1497–1505 (1995). [CrossRef]  

24. Y. J. Lee, “Determination of 3D molecular orientation by concurrent polarization analysis of multiple Raman modes in broadband CARS spectroscopy,” Opt. Express 23, 29279 (2015). [CrossRef]   [PubMed]  

25. A. Cunningham, G. Davies, and I. Ward, “Determination of molecular orientation by polarized infra-red radiation in an oriented polymer of high polarizability,” Polymer 15, 743–748 (1974). [CrossRef]  

26. H. Nakahara and K. Fukuda, “Studies on molecular orientation in multilayers of long-chain anthraquinone derivatives by polarized infrared spectra,” J. Colloid Interface Sci. 69, 24–33 (1979). [CrossRef]  

27. F. Hide, N. A. Clark, K. Nito, A. Yasuda, and D. M. Walba, “Dynamic Polarized Infrared Spectroscopy of Electric Field-Induced Molecular Reorientation in a Chiral Smectic- A Liquid Crystal,” Phys. Rev. Lett. 75, 2344–2347 (1995). [CrossRef]   [PubMed]  

28. G. J. Simpson, S. G. Westerbuhr, and K. L. Rowlen, “Molecular Orientation and Angular Distribution Probed by Angle-Resolved Absorbance and Second Harmonic Generation,” Anal. Chem. 72, 887–898 (2000). [CrossRef]   [PubMed]  

29. S. G. Boxer, A. Kuki, K. a. Wright, B. a. Katz, and N. H. Xuong, “Oriented properties of the chlorophylls: Electronic absorption spectroscopy of orthorhombic pyrochlorophyllide a-apomyoglobin single crystals,” Proc. Natl. Acad. Sci. United States Am. 79, 1121–1125 (1982). [CrossRef]  

30. Y. Nagasaki, T. Yoshihara, and Y. Ozaki, “Polarized Infrared Spectroscopic Study on Hindered Rotation around the Molecular Axis in the Smectic-C* Phase of a Ferroelectric Liquid Crystal with a Naphthalene Ring. Application of Two-Dimensional Correlation Spectroscopy to Polarization Angle-Dependen,” The J. Phys. Chem. B 104, 2846–2852 (2000). [CrossRef]  

31. P. Papadopoulos, J. Sölter, and F. Kremer, “Structure-property relationships in major ampullate spider silk as deduced from polarized FTIR spectroscopy,” Eur. Phys. J. E 24, 193–199 (2007). [CrossRef]   [PubMed]  

32. C. W. Myers and S. L. Cooper, “Analysis of some errors in the calculation of orientation functions for polymers from infrared dichroism measurements,” Appl. Spectrosc. 48, 72 (1994). [CrossRef]  

33. T. G. Mayerhöfer, H. Mutschke, and J. Popp, “Employing Theories Far beyond Their Limits-The Case of the (Boguer-) Beer-Lambert Law,” ChemPhysChem. 17, 1948–1955 (2016). [CrossRef]   [PubMed]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Polarization angle dependence of transmitted light intensity through a collection of aligned transition dipole moments. The incident light, I(0), propagating along the z axis, is linearly polarized with the polarization angle rotated by η from the x axis in the xy plane. All transition dipole moments are aligned parallel in the xz plane with the tilting angle of θ from the z axis. The magnitude of the transition dipole moment, equivalent to the absorption cross section, is represented by ε. The terms d and c denote the sample thickness and the concentration, respectively. As the light passes through the sample, only Ex is reduced by absorption but Ey is unaffected.
Fig. 2
Fig. 2 Schematic presentation of broadening and rotation of molecular orientation. The orientational broadening is represented by ρ(β), where β denotes the axial angle of an individual μ from the mean orientation direction, 〈μ〉. The molecular rotation is represented by the rotation matrix R(ψ, θ, ϕ) from the molecular coordinate to the laboratory coordinate. The incident light propagates in the direction of the z axis, and the light polarization plane is on the xy plane. η is the polarization angle with respect to the x axis.
Fig. 3
Fig. 3 (a) Plots of the von Mises-Fisher function ρvMF(β) for two different κ values. (b) and (c) Plots of absorptance calculated from Eqs. (10) and (12) for the primary (red) and the secondary (blue) transition dipole moments using two different sets of orientation angles of the mean orientation directions. The 3D angles used for calculation are indicated above each figure. For comparison, absorptance profiles from unbroadened ODF (κ = ∞) are plotted as the solid lines. The dashed lines indicate absorptance profiles calculated from broadened ρvMF(β) with κ = 10 (b) and κ = 50 (c). For both (b) and (c), α 1 ° = 0.9 and α 2 ° = 0.9.
Fig. 4
Fig. 4 A flow chart of the concurrent polarization analysis method for determination of the 3D angles (ψ, θ, and ϕ) of the mean molecular orientation and the order parameter 〈P2〉 of the local orientation distribution using polarization profiles of two orthogonal IR modes. The intermediate parameters (Δη, M, and N) are calculated from the six input observables of the phase angles and the maximum and minimum absorptances of the two polarization profiles.
Fig. 5
Fig. 5 (a), (b), (c) Comparison of polarization angle-dependent absorptance calculated for ODF-broadened orientation based on the simplified and the original forms of α(η). The calculation is performed for the primary mode, and the axial angle θ = 60°. The von Mises-Fisher function with κ = 10 is used as the model ODF for the numerical calculation. (d), (e) The maximum-to-minimum ratio, represented by M in Eq. (27) is plotted as a function of α° for different θ.

Tables (1)

Tables Icon

Table 1 Demonstration of determination of the 3D angles and 〈P2〉 using observables and intermediate parameters from the simulated polarization absorptance profiles shown in Fig. 3. The input uncertainties for phase angle η and absorptance α are assumed to be 2° and 5%, respectively, for Fig. 3b, and 5° and 10% for Fig. 3c. The parentheses for θ and ϕ indicate the second solution for the 3D angles of the mirror image molecular orientation with respect to the polarization plane.

Equations (41)

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

I ( d ) = I ( 0 ) { 1 [ 1 exp ( 2 ε c d sin 2 θ ) ] cos 2 ( η ψ ) }
T ( η ) I ( d ) / I ( 0 ) = 1 [ 1 exp ( 2 ε c d sin 2 θ ) ] cos 2 ( η ψ )
A ( η ) = log { 1 [ 1 exp ( 2 ε c d sin 2 θ ) ] cos 2 ( η ψ ) }
α ( η ) = [ 1 exp ( 2 ε c d sin 2 θ ) ] cos 2 ( η ψ )
α ( η ) = { 1 [ 1 ( 2 ε c d sin 2 θ ) + ( 2 ε c d sin 2 θ ) 2 / 2 ] } cos 2 ( η ψ ) = [ ( 2 ε c d sin 2 θ ) ( 2 ε c d sin 2 θ ) 2 / 2 + ] cos 2 ( η ψ ) .
α ( η ) ( 2 ε c d ) sin 2 θ cos 2 ( η ψ ) α ° sin 2 θ cos 2 ( η ψ )
α ( η ) = α ° [ e ^ E ( η ) e ^ μ ( ψ , θ ) ] 2
R ( ψ , θ , ϕ ) = ( cos ψ sin ψ 0 sin ψ cos ψ 0 0 0 1 ) ( cos θ 0 sin θ 0 1 0 sin θ 0 cos θ ) ( cos ϕ sin ϕ 0 sin ϕ cos ϕ 0 0 0 1 )
α 1 ( η ) = α 1 ° 0 2 π 0 π [ e ^ E ( η ) e ^ μ 1 ] 2 ρ ( β ) sin β d β d γ = α 1 ° 0 2 π 0 π [ e ^ E ( η ) R ( ψ , θ , ϕ ) R ( γ , β , 0 ) e ^ z ] 2 ρ ( β ) sin β d β d γ
α 1 ( η ) α 1 ° = 1 2 [ 1 sin 2 θ cos 2 ( η ψ ) ] + 1 2 [ 3 sin 2 θ cos 2 ( η ψ ) 1 ] 0 π cos 2 β ρ ( β ) sin β d β
α 2 ( η ) = α 2 ° 0 2 π 0 π [ e ^ E ( η ) R ( ψ , θ , ϕ ) R y ( π 2 ) R ( γ , β , 0 ) e ^ z ] 2 ρ ( β ) sin β d β d γ
α 2 ( η ) α 2 ° = 1 2 { [ cos θ cos ϕ cos ( η ψ ) + sin ϕ sin ( η ψ ) ] 2 1 } + 1 2 { 3 [ cos θ cos ϕ cos ( η ψ ) + sin ϕ sin ( η ψ ) ] 2 1 } × 0 π cos 2 β ρ ( β ) sin β d β
ρ ( β ) = n = 0 ( n + 1 2 ) P n P n ( cos β ) .
P 0 ( cos β ) = 1
P 2 ( cos β ) = 1 2 ( 3 cos 2 β 1 ) .
0 π cos 2 β ρ ( β ) sin β d β = 2 P 2 + 1 3 .
α 1 ( η ) α 1 ° = P 2 sin 2 θ cos 2 ( η ψ ) + 1 3 ( 1 P 2 ) .
α 2 ( η ) α 2 ° = P 2 [ cos θ cos ϕ cos ( η ψ ) + sin ϕ sin ( η ψ ) ] 2 + 1 3 ( 1 P 2 ) .
P 2 = α max α min α max + 2 α min = P 2 sin 2 θ 1 P 2 cos 2 θ
ψ = η 1 , max ( 0 ψ π )
cos ϕ cos θ sin Δ η = sin ϕ cos Δ η
Δ η = tan 1 ( tan ϕ sec θ )
α 1 , min α 1 ° = 1 3 ( 1 P 2 )
α 1 , max α 1 ° = P 2 sin 2 θ + 1 3 ( 1 P 2 )
α 2 , min α 2 ° = 1 3 ( 1 P 2 )
α 2 , max α 2 ° = P 2 ( cos 2 ϕ cos 2 θ + sin 2 ϕ ) + 1 3 ( 1 P 2 )
M ( α 1 , max α 1 , min 1 ) = 3 P 2 1 P 2 × sin 2 θ
N ( α 2 , max α 2 , min 1 ) = 3 P 2 1 P 2 ( cos 2 ϕ cos 2 θ + sin 2 ϕ )
P 3 P 2 1 P 2
sin 2 θ = M / P
cos 2 ϕ = ( P N ) / M
( N / P ) cos 2 Δ η = cos 2 ϕ cos 2 θ = cos 2 ϕ ( 1 sin 2 θ )
( P M ) ( P N ) = M N cos 2 Δ η
P = 1 2 [ ( M + N ) ± ( M + N ) 2 4 M N ( 1 cos 2 Δ η ) ] = 1 2 [ ( M + N ) ± ( M N ) 2 + 4 M N cos 2 Δ η ]
cos 2 ϕ = 1 2 M [ ( M N ) ± ( M N ) 2 + 4 M N cos 2 Δ η ] 0
cos 2 ϕ = 1 2 M [ ( M N ) + ( M N ) 2 + 4 M N cos 2 Δ η ]
P = 1 2 [ ( M + N ) + ( M N ) 2 + 4 M N cos 2 Δ η ]
P 2 = P P + 3
θ = sin 1 M / P ( 0 θ π )
ϕ = cos 1 [ ± ( P N ) / M ] ( π ϕ π )
δ f ( x 1 , , x n ) = ( f x 1 ) 2 ( δ x 1 ) 2 + + ( f x n ) 2 ( δ x n ) 2 .
Select as filters


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