Open Access
19 April 2012 Second harmonic generation signal in collagen fibers: role of polarization, numerical aperture, and wavelength
Oscar del Barco, Juan M. Bueno
Author Affiliations +
Abstract
The spatial distribution of second harmonic generation (SHG) signal from collagen fibers for incident elliptical polarized light has been modeled. The beam was assumed to focus on a horizontal fiber through a microscope objective. For elliptical polarized states located along a vertical meridian of the Poincare sphere, the SHG intensity has been optimized in terms of the incident wavelength and the numerical aperture (NA) of the objective. Our results show that polarization modulates the SHG signal. Elliptical polarization can generate high signals (even greater than those corresponding to linear polarization) when combined with appropriate values of both incident wavelength and NA. On the other hand, the SHG intensity might also be identically zero for particular elliptical polarization states, a condition that depends exclusively on the ratio of hyperpolarizabilities of the collagen fibers. The highest forward-to-backward SHG signal distribution occurs along the propagating direction, depends on the incident wavelength, and reduces when NA increases. Furthermore, the direction of maximum SHG emission was found to be more sensitive to changes in the NA rather than variations in the incident wavelength. These findings could help to optimize the experimental conditions of multiphoton microscopes and to increase SHG signals from biological tissues containing collagen.

1.

Introduction

During the last years there has been an increasing interest in the use of multiphoton microscopy for biomedical applications. In particular, second harmonic generation (SHG) imaging is a noninvasive technique that conserves energy, provides useful information on microscopic structures, and allows optical three-dimensional (3-D) sectioning and deep penetration.1,2 Although, it is well known that SHG signal originates from noncentrosymmetric structures such as collagen,1,3 this nonlinear process is also sensitive to local anisotropy,4 distribution of molecular hyperpolarizabilities,5 and phase-matching properties.6 SHG is a coherent process involving phase-matching considerations whose emission is mainly directed rather than isotropic. SHG directionality depends on the distribution of induced dipoles within the focal volume where the nonlinear process takes place.7 Since momentum conservation provides a preferred SHG signal in the forward direction (F-SHG) (i.e., along the incident direction of the laser beam), most nonlinear microscopes adopt this experimental configuration. Clinical applications involving SHG imaging of collagen structures in animals often require a backward collection geometry (B-SHG). Moreover, the structure of extracellular collagen plays an important role in cancerous processes (tumor development), aging and wound healing.8 Several works have compared F-SHG and B-SHG imaging in different specimens containing collagen, such as endometrium tissue,4 tendons,9,10 corneal stroma,11,12 skin,6,13 and tumors,14 among others. All these studies conclude that the SHG imaged structures strongly depend on the recording configurations of the microscope. On the other hand, theoretical works discussing the differences between F-SHG and B-SHG signals have also been reported.10,15

Linear polarized light has been shown to visualize different features of samples containing collagen, when imaged by means of SHG microscopy.1620 Whereas, some experiments and theoretical models were restricted to study the effects of incident linear polarized light on SHG imaging,10,15,21 others have investigated the use of elliptical polarization to monitor molecular orientation22 and to analyze the SHG signal from surfaces and films.23,24 Some studies have also reported the effects of circular and elliptical polarization on the SHG intensity.25,26 In particular, Schön et al.26 showed that the dichroic mirrors used in the microscope collection optics can induce high levels of ellipticity in the incident beam, what leads to severe distortions in the polarization response. These distortions can be corrected by combining a half-wave plate and a quarter-wave plate.25

The aim of this work is to model and characterize the spatial distribution of SHG signal originated from collagen for focused elliptical polarized light. In Sec. 2 this theoretical model is developed. As a result, an analytical expression for the SHG intensity signal I2ω as a function of the incident state of polarization, the fundamental wavelength λω, and the numerical aperture (NA) of the microscope objective is derived. An extensively numerical analysis of I2ω is described in Sec. 3, while Sec. 4 contains the discussion and the main conclusions of this work. Moreover, two appendixes with details on the calculation of the dipole moment induced by the incident electric field (Appendix A) as well as the total radiated SHG signal in the far field approximation (Appendix B) are also included.

2.

SHG Intensity Signal for Elliptically Polarized Light Focused on a Collagen Fiber

Let us consider a beam of elliptical polarized light with fundamental frequency ω propagating in the z^ direction (see Fig. 1) focusing through a microscope objective with numerical aperture NA=nωsinΘ on a collagen fiber lying along the x^ direction. Θ corresponds to the maximum angle at which the light rays enter the objective, while nω represents the index of refraction of collagen. The dependence of nω with the fundamental wavelength λω of the incident light has been reported to be27

Eq. (1)

nω(λω)=1426+19476λω21131066900λω4,
where λω is expressed in nanometers. The perpendicular components of the elliptical polarized light Eω,x and Eω,y can be written as

Eq. (2)

Eω,x(x,y,z)=iEω,x(0)exp[(x2+y2wxy2+z2wz2)]exp[iξkωz],

Eq. (3)

Eω,y(x,y,z)=iEω,y(0)exp[(x2+y2wxy2+z2wz2)]exp[i(ξkωz+δ)],
where δ is the phase difference between these components, and kω=(2πnω)/λω the wave number in a medium with index of refraction nω. ξ corresponds to the wave-vector reduction factor that characterizes the phase shift experienced by a Gaussian beam in the vicinity of a focal center.15,28 For NAs lower than 1.2, ξ can be approximated by cos(Θ/2). On the other hand, wxy and wz represent the 1/e radii of the focal ellipse in the lateral and axial directions, respectively8,10,29

Eq. (4)

wxy=0.32λωnωsinΘwz=0.53λωnω(1cosΘ).

Fig. 1

A beam of elliptical polarized light with fundamental frequency ω, propagating in the z^ direction, focusing through a microscope objective with numerical aperture NA on a collagen fiber lying along the x^ direction.

JBO_17_4_045005_f001.png

As shown in Fig. 1, ψ symbolizes the ellipticity of the incident beam. For the sake of simplicity, the azimuthal angle of the ellipse of polarization has been chosen to be zero. This light conditions represent all the possible polarization states located along the null azimuth meridian on the Poincare sphere, what is consistent with sets of experimental polarization states provided by liquid-crystals.30 This assumption will notably simplify our theoretical model since each component of the incident field Eω,x and Eω,y induces an SHG dipole moment along the collagen fiber (x^ direction). Appendix A explicitly shows that the dipole moments μ2ω,x(x) and μ2ω,x(y) induced by the electric fields Eω,x and Eω,y can , respectively be expressed as

Eq. (5)

μ2ω,x(x)(x,y,z)=12Eω,x2βxxx,

Eq. (6)

μ2ω,x(y)(x,y,z)=12Eω,y2βxyy,
where βxxx and βxyy correspond to the first hyperpolarizabilities contributing to the SHG field. Adding both quantities, the expression for the total SHG dipole moment will be

Eq. (7)

μ2ω(x,y,z)=μ2ω,x(x)(x,y,z)+μ2ω,x(y)(x,y,z)=12(Eω,x2βxxx+Eω,y2βxyy).

This SHG dipole moment is directly related to the SHG far field E2ω radiated in the (θ,φ) direction15,28,31,32

Eq. (8)

E2ω(Ψ)=ημ2ωsin(Ψ)exp[ik2ω·r]Ψ^,
where η=ω2/πϵ0c2 and Ψ^ represents the angle between the x^ axis of the collagen fiber and the direction of emission r of the SHG field. ϵ0 is the free-space permittivity, c the vacuum speed of light, and k2ω=(2πn2ω)/λ2ω the wave number for the SHG signal.

The total radiated SHG signal in the far field approximation is obtained by integration from all scatterers that have a dipole volume density defined by Nv15 (see Appendix B)

Eq. (9)

E2ω,x(r,θ,φ)=12[(π2)3wxy2wz]Nvηr(sin2θsin2φ+cos2θ)1/2×exp[k2ω28(wxy2sin2θ+wz2(cosθξ)2)]×[(Eω,x(0))2βxxx+(Eω,y(0))2βxyyexp[i2δ]],
with ξ=ξ(nω/n2ω).

Therefore, the SHG intensity signal I2ω can be written in terms of the electric field as

Eq. (10)

I2ω(r,θ,φ)=12n2ωϵ0c|E2ω,x|2=[18n2ωϵ0c(NvV)2](ηr)2(sin2θsin2φ+cos2θ)×exp[k2ω24(wxy2sin2θ+wz2(cosθξ)2)]×[(Eω,x(0))4βxxx2+(Eω,y(0))4βxyy2+2(Eω,x(0))2(Eω,y(0))2βxxxβxyycos(2δ)],
where V is the active SHG volume31

Eq. (11)

V=(π2)3wxy2wz.
From Fig. 1 the perpendicular components of the incident amplitude Eω(0) can easily be expressed as

Eq. (12)

Eω,x(0)=Eω(0)cosψEω,y(0)=Eω(0)sinψ,
then, Eq. (10) reduces to

Eq. (13)

I2ω(ψ,r,θ,φ)=[18n2ωϵ0c(NvV)2](ηr)2(sin2θsin2φ+cos2θ)×exp{k2ω24[wxy2sin2θ+wz2(cosθξ)2]}×|Eω(0)|4[cos4ψβxxx2+sin4ψβxyy2+12sin2(2ψ)cos(2δ)βxxxβxyy].

Equation (13) shows how the SHG signal explicitly depends on the ellipticity of the incident beam, ψ. Since the phase δ is equal to ±π/2 for the elliptical polarized light here used (zero azimuth),30 the final expression for the SHG intensity signal will be

Eq. (14)

I2ω(ψ,r,θ,φ)=[Iω22n2ωnω2ϵ0c(NvV)2](ηr)2(sin2θsin2φ+cos2θ)×exp{k2ω24[wxy2sin2θ+wz2(cosθξ)2]}×[cos4ψβxxx2+sin4ψβxyy212sin2(2ψ)βxxxβxyy],
where Iω=(1/2)nωϵ0c|Eω(0)|2 corresponds to the intensity of the incident beam.

It can be noticed that the last term in Eq. (14) contains the polarization contribution to I2ω via the ellipticity of the incident beam ψ. This factor must be taken into consideration for a complete description of the SHG intensity, as it will be shown in detail in the next section.

3.

Analysis of the SHG Intensity Signal

Once our general expression of the SHG intensity signal I2ω for elliptically polarized light has been obtained, a numerical analysis of this magnitude as a function of different parameters, such as the polarization state of the incident beam, the incident wavelength λω, and the NA of the microscope objective will be carried out.

3.1.

Polarization Dependence of I2ω

As above noticed, the analytical expression for I2ω [Eq. (14)] includes a factor involving the polarization state of the incident light (elliptically polarized in general). Figure 2 shows the strong dependence of the SHG intensity on the ellipticity ψ of the incident light. Both F-SHG I2ω(F) (i.e., π/2<θ<π/2) and B-SHG I2ω(B) (i.e., π/2<θ<3π/2) signals are plotted versus the polar coordinates (θ,φ) for four different polarization states (linear horizontal (H), right-handed elliptical with ellipticity 30 deg (ER,30deg), right-handed circular (CR) and linear vertical (V). Since there exist several orders of magnitude between F-SHG and B-SHG, I2ω(B) has been plotted in logarithmic scale, for the sake of clarity. For these plots, λω and NA were chosen to be 700 nm and 1.2, respectively. The refractive indexes corresponding to the SHG signal n2ω and the incident light nω were computed by using Eq. (1) (1.509 and 1.461, respectively). The term {[Iω2n2ω(NvV)2]/(2nω2ϵ0c)}×(η/r)2 was normalized to 1 for simplicity, and the selected value of the ratio of hyperpolarizabilities ρ=βxxx/βxyy was 2.6 [ρ has been reported to range between 3 and 3 (Ref. 33)]. It can be observed that the maximum F-SHG intensity value occurs at the polar angle θmax=arccos[ξ(nω/n2ω)] where the phase match condition k2ωcosθ2ξkω=0 is satisfied.10,15 For both F-SHG and B-SHG components, the SHG signal is maximum when the incident-polarized light is linear horizontal H (i.e., parallel to the collagen fibers’ direction). Unlike expected, vertical polarization V does not minimize the SHG intensity. Right-handed circular polarized light (CR) makes the SHG signal even smaller. This will be analyzed and discussed in the following paragraphs.

Fig. 2

F-SHG (I2ω(F), left column) and B-SHG (I2ω(B), right column) intensities versus polar coordinates (θ,φ) for four different polarization states (linear horizontal (H), right-handed elliptical with ellipticity 30 deg (ER,30deg), right-handed circular (CR) and linear vertical (V). The fundamental wavelength λω and the NA of the microscope objective have been chosen to be 700 nm and 1.2, respectively.

JBO_17_4_045005_f002.png

Let us now consider all the possible polarization states located along the meridian with null azimuth on the Poincare sphere. In this sense, Fig. 3(a) presents the maximum F-SHG intensity values I2ω,max(F) as a function of the ellipticity of the incident beam ψ. The four polarization states described in Fig. 2 have also been labeled as well as the left-handed circular polarization state (CL). In particular, when ψ equals 58.2 or 121.8 deg, the SHG intensity I2ω,max(F) is identically zero, a fact that reveals the importance of the incident polarization state on the SHG signal. As expected, H polarized light yields a maximum I2ω,max(F) value (see again Fig. 2). This SHG intensity is reduced when moving along the selected meridian on the Poincare sphere.30 Figure 3(b) depicts the corresponding maximum values for F-SHG and B-SHG signals in logarithmic scale for direct comparison. It is observed that both curves have similar shapes with different orders of magnitude.

Fig. 3

(a) Maximum F-SHG intensity values I2ω,max(F) and (b) maximum F-SHG and B-SHG intensity signals as a function of the ellipticity of the incident light ψ. Logarithmic scale has been used in Fig. 3(b) for a better comparison. CL indicates left-handed circular polarized light.

JBO_17_4_045005_f003.png

After analyzing the results shown in Fig. 3, a question arises: could I2ω become identically zero by solely modifying the polarization state of the incident light? In order to answer this question, the roots of Eq. (14) were analytically computed. I2ω vanishes for the following values of the ellipticity ψ0

Eq. (15)

ψ0=arccos(±1/1+ρ).
This expression depends exclusively on the ratio of hyperpolarizabilities ρ. Accordingly, Fig. 4 plots the values of ψ0 as a function of ρ calculated via Eq. (15). It can be observed that, for a fixed value of ρ, there exist two ellipticity values (or equivalently, two polarization states in the meridian with null azimuth of the Poincare sphere) that vanish the SHG intensity. The existence of such polarization states is limited exclusively to values of ρ>1, as can easily be deduced from simple inspection of Eq. (15).

Fig. 4

Ellipticity values ψ0 versus the ratio of hyperpolarizabilities ρ calculated via Eq. (15). For a fixed ρ there exist two possible values of ψ0, which correspond to two different states along the meridian of the Poincare sphere with azimuth zero.

JBO_17_4_045005_f004.png

3.2.

Effects of the Fundamental Wavelength and the Numerical Aperture on I2ω

Results shown in Figs. 2Fig. 34 have been derived for the parameters λω=700nm and NA=1.2. However, it is also interesting to investigate how the incident wavelength and the NA affect the SHG signal for different incident polarization states. In particular, Fig. 5(a) and 5(b) present two polar diagrams where the B-SHG signal has been plotted as a function of 2ψ for different values of the λω and NA, respectively. For a better comparison, the curves have been normalized to the maximum B-SHG intensity value (λω=1000nm and NA=1.2 for the present case). As expected, for fixed values of the wavelength and the NA, the maximum B-SHG intensity occurs for horizontal polarized light H (what agrees with Figs. 2 and 3). However, for a chosen NA value, I2ω(B) increases with λω, as shown in Fig. 5(a). That is, the maximum B-SHG intensity value will strongly depend on the incident wavelength value. As an example, an elliptical polarized light with 2ψ=15° and λω=850nm focusing on a microscope objective with NA=1.2, will generate a B-SHG signal 1.4 times larger than the signal produced by a horizontal polarized light with λω=700nm. From Fig. 5(b), the elliptical polarized light above cited will also generate a B-SHG signal 3.4 times larger than the corresponding H polarized light with an NA of 1.0. As a direct consequence, optimized values of the B-SHG signal can be achieved by using elliptical light combined with appropriate values of NA and λω. Similar results have been found for the F-SHG intensity signal (not included in this work) so, these conclusions can also be extended to this SHG collection geometry.

Fig. 5

B-SHG intensity I2ω,max(B) as a function of 2ψ for different values of (a) the incident wavelength λω (700, 850, and 1000 nm) and (b) the numerical aperture NA (1.0, 1.1, and 1.2). The results have been normalized to the maximum value of the B-SHG intensity (λω=1000nm and NA=1.2, in our case) for a better comparison.

JBO_17_4_045005_f005.png

On the other hand, from Eq. (14) it can easily be noticed that the F/B SHG ratio I2ω(F)/I2ω(B) does not depend on the polarization state of the incident light (the last term cancels out). So, the effects of the incident wavelength λω (i.e., chromatic dispersion) on the F/B SHG ratio have been investigated. Figure 6 shows the spatial distribution (expressed in polar coordinates (θ,φ)) of the F/B ratio for three different λω values. For this plot, NA and ρ were chosen to be 1.2 and 2.6, respectively. It can be observed that the maximum F/B SHG ratio occurs along the propagating direction z^, that is, for θ=0. Moreover, the comparison of the F/B SHG spatial distributions for different λω shows an increase of several orders of magnitude for higher values of λω.

Fig. 6

F/B SHG intensity ratio I2ω(F)/I2ω(B) versus the polar coordinates (θ,φ) for three different values of the fundamental wavelength λω (grey: 700 nm, dark grey: 850 nm, and black: 1000 nm).

JBO_17_4_045005_f006.png

The effects of the NA on the F/B SHG ratio are presented in Fig. 7. The plot represents the I2ω(F)/I2ω(B) spatial distribution for three different NA values. In this case, λω=850nm and ρ=2.6. It can be observed that the NA of the microscope objective plays a fundamental role in the detected SHG signal. In particular, as the NA increases the spatial distribution spreads, and the F/B SHG ratio (which is always maximum along the direction of light propagation) decreases several orders of magnitude.

Fig. 7

Spatial distribution of the F/B SHG intensity ratio for different numerical aperture values (grey: 1.2, dark grey: 1.1, and black: 1.0).

JBO_17_4_045005_f007.png

It is also worth while to study how the angle of maximum SHG intensity θmax=arccos[ξ(nω/n2ω)] (which is independent on the ratio of hyperpolarizabilities ρ) depends on both λω and NA. In this sense, Fig. 8(a) plots the polar angle θmax as a function of the NA for λω=850nm. These results show that, although θmax occurs always at forward propagation angles (i.e., π/2<θ<π/2), it separates from the propagating direction z^ as the NA increases. In Fig. 8(b) this polar angle is represented versus the incident wavelength for NA=1.2. We can conclude that the influence of the NA is stronger than the dispersion effects due to changes in the fundamental wavelength λω. In particular, for the conditions here considered, when NA varied from 0.9 to 1.2, the angle of maximum SHG intensity θmax increased 65%. However this angle only increased 4% when λω moved from 700 to 1000 nm.

Fig. 8

Angle of maximum SHG intensity signal θmax versus (a) the numerical aperture NA (λω=850nm) and (b) the incident wavelength λω (NA=1.2).

JBO_17_4_045005_f008.png

4.

Discussion and Conclusions

A theoretical model to calculate the spatial distribution of I2ω in collagen fibers for focused elliptical polarized light has been developed. A general expression involving the polar coordinates (θ,φ), the incident wavelength λω, the NA of the microscope objective, and the ellipticity of the incident beam ψ has been obtained. For the sake of simplicity, collagen fibers were considered lying along the x^ direction. Under this assumption the spatial distribution was found to be anisotropic (see Fig. 2) for both F-SHG and B-SHG signals, although, the modulation of this anisotropy strongly depended on the incident polarization state. For fixed values of λω and NA, the SHG signal presented maximum values for H linear polarized light (i.e., parallel to the direction of the collagen fibers) as shown in Figs. 2 and 3. This result is consistent with previously published work on SHG intensity signal for linear polarized light.15 Unlike expected, vertical polarization does not minimize the SHG intensity, since CR light makes this magnitude even smaller (see Fig. 3). However, I2ω might also be identically zero for particular values of the ellipticity ψ0 of the incident polarized light [see Eq. (15)]. This occurs when ψ0=arccos(±1/1+ρ), a condition that depends exclusively on the ratio of hyperpolarizabilities ρ (Fig. 4). This fact must be taken into account when optimizing the SHG signal originated from a particular collagen sample. Incident polarization states must be “far away” from those verifying this condition. The effect of λω and the NA of the microscope objective on the SHG intensity, besides its dependence on the incident polarization state, has also been investigated in detail. In general, for a fixed polarization state, the higher the values of the incident wavelength and the NA, the higher the SHG signal. Moreover, elliptical polarized light can produce SHG signals higher than those corresponding to H polarization when combined with appropriate values of both λω and NA (see Fig. 5).

Multiphoton microscopy can be used in both F and B configurations. It has been experimentally shown that SHG microscopy images of tissues containing collagen are seen differently depending on the collection geometry used (forward or backward).10,11,17 Having this into account, an in-depth analysis of the spatial distribution of the F/B SHG ratio has also been carried out. In particular, this distribution presents a maximum value along the propagating direction (θ=0) and significantly increases with the fundamental wavelength λω (Fig. 6). This is an important fact that has previously been addressed by different authors.12,34 On the other hand, as the NA increases, the spatial distribution spreads and the F/B SHG ratio decreases several orders of magnitude, as shown in Fig. 7. This corroborates the fundamental role played by the NA of the microscope objective in SHG detection.

Stoller et al. reported a study on the dependence of the F-SHG peak with the incident wavelength.35 They obtained an increase of the normalized SHG peak signal with the NA, without taking into account the spatial resolved dependence. In a theoretical study, Chang and co-workers explored the combination effect of incident linear polarized light and the NA15. They reported that for NA<0.4, the total SHG signal did not depend on polarization. However for higher values of NA there is a nonlinear dependence, which is particular for each polarization state. Since the present work explores the F/B SHG, the influence of the polarization state cancels out, as can easily deduced from Eq. (14).

Our numerical results have finally showed that independently of λω and NA values, the angle of maximum SHG intensity θmax always occurs at forward propagation angles. This will benefit the microscopes with forward experimental configurations. Furthermore, the influence of the NA on this angle is stronger than the dispersion effects due to changes in the fundamental wavelength (see Fig. 8).

SHG imaging of tissues composed of collagen is a topic of great interest in Biomedicine. As already reported, polarization dependence of the SHG signal can be used to increase the contrast in collagen SHG imaging,16 to obtain information unreachable with intensity-based SHG microscopy17,36 or to explore the sources of SHG signal within the specimen,20,21 among others. The analysis of biological collagen-made structures with a noninvasive tool is important to understand and/or to monitor the effects of injury, healing process, surgery, pathologies, and thermal damages. The results reported here provide detailed information on how polarization can be used to optimize the SHG signal. Moreover, the analysis of the effects of both NA and incident wavelength might also be used to maximize the signal recorded with a multiphoton microscope. In particular, the implementation of these findings into SHG imaging devices would help to reduce the power of the incident laser beam in order to minimize photodamage, which is an issue in experiments with both ex-vivo and living tissues.

Appendices

Appendix A:

Dipole Moment Induced by the Incident Electric Field

In the particular case of a linear polarized beam at an angle α with respect to the collagen fiber (x^ direction), the SHG dipole moment can be written as15

Eq. (16)

μ2ω(x,y,z)=12Eω2(x,y,z)×(βxxxcos2α+βxyysin2αβxyysin2α0).

Taking into account that the component Eω,x of the incident field is linear polarized with α=0, Eq. (16) yields

Eq. (17)

μ2ω(x)(x,y,z)=12Eω,x2(x,y,z)×(βxxx00).

This expression reveals the absence of SHG dipole components in y^ and z^ directions. For this reason, the vectorial character of μ2ω(x) and Eω,x can be omitted and the SHG dipole moment induced by the electric field Eω,x can be written as [see Eq. (5)]

Eq. (18)

μ2ω(x)(x,y,z)=12Eω,x2(x,y,z)βxxx.

Equivalently, the component of the incident field Eω,y is also linear polarized with α=π/2, so Eq. (16) turns into

Eq. (19)

μ2ω(y)(x,y,z)=12Eω,y2(x,y,z)×(βxyy00),
and the SHG dipole moment induced by Eω,y can be expressed as [see Eq. (6)]

Eq. (20)

μ2ω(y)(x,y,z)=12Eω,y2(x,y,z)βxyy.

Appendix B:

Total Radiated SHG Signal in the Far Field Approximation

The SHG far field is given by Eq. (8)

Eq. (21)

E2ω(Ψ)=ημ2ωsin(Ψ)exp[ik2ω·r]Ψ^,
where, as mentioned, η=ω2/πϵ0c2, Ψ represents the angle between the x^ axis and the emission direction r of the SHG field and sin(Ψ)=(sin2θsin2φ+cos2θ)1/2. The total radiated SHG signal in the far field related to the SHG dipole moment μ2ω(x) is the integrated one from all scatterers that have the dipole volume density defined by Nv. In polar coordinates, E2ω,x(x)(θ,φ) can be expressed as15

Eq. (22)

E2ω,x(x)(θ,φ)=ηNvr(sin2θsin2φ+cos2θ)1/2dxdydz12(Eω,x(0))2βxxx×exp(2x2+y2wxy22z2wz2+2iξkωz)×exp[ik2ω(zcosθ+ysinθsinφ+xsinθcosφ)],
where Eq. (2) has been used for the calculation of the incident electric field Eω,x. Upon integration of Eq. (22) in x, y, and z directions from negative infinite to positive infinite, the total radiated SHG signal E2ω,x(x), is given by15

Eq. (23)

E2ω,x(x)(θ,φ)=12[(π2)3wxy2wz]Nvηr(sin2θsin2φ+cos2θ)1/2×exp[k2ω28(wxy2sin2θ+wz2(cosθξ)2)](Eω,x(0))2βxxx,
where ξ=ξ(nω/n2ω).

Similarly, via Eq. (3), the following expression for the total radiated SHG signal in the far field related to the SHG dipole moment μ2ω(y) can be obtained

Eq. (24)

E2ω,x(y)(θ,φ)=12[(π2)3wxy2wz]Nvηr(sin2θsin2φ+cos2θ)1/2×exp[k2ω28(wxy2sin2θ+wz2(cosθξ)2)](Eω,y(0))2βxyyexp[i2δ].
Finally, the expression for the total SHG signal E2ω,x is obtained by adding Eq. (23) and (24)

Eq. (25)

E2ω,x(θ,φ)=12[(π2)3wxy2wz]Nvηr(sin2θsin2φ+cos2θ)1/2×exp[k2ω28(wxy2sin2θ+wz2(cosθξ)2)]×[(Eω,x(0))2βxxx+(Eω,y(0))2βxyyexp[i2δ]].

Acknowledgments

This work has been supported by “Ministerio de Ciencia e Innovación,” Spain (Grants FIS2010-14926 and CONSOLIDER-INGENIO 2010, CSD2007-00033 SAUUL); Fundación Séneca (Región de Murcia, Spain), Grant 4524/GERM/06.

References

1. 

Y. Guoet al., “Second-harmonic tomography of tissues,” Opt. Lett., 22 (17), 1323 –1325 (1997). http://dx.doi.org/10.1364/OL.22.001323 OPLEDP 0146-9592 Google Scholar

2. 

P. J. Campagnolaet al., “High-resolution nonlinear optical imaging of live cells by second harmonic generation,” Biophys. J., 77 (6), 3341 –3349 (1999). http://dx.doi.org/10.1016/S0006-3495(99)77165-1 BIOJAU 0006-3495 Google Scholar

3. 

S. FineW. P. Hansen, “Optical second harmonic generation in biological systems,” Appl. Opt., 10 (10), 2350 –2353 (1971). http://dx.doi.org/10.1364/AO.10.002350 APOPAI 0003-6935 Google Scholar

4. 

G. Coxet al., “3-D imaging of collagen using second harmonic generation,” J. Struct. Biol., 141 (1), 53 –62 (2003). http://dx.doi.org/10.1016/S1047-8477(02)00576-2 JSBIEM 1047-8477 Google Scholar

5. 

I. FreundM. Deutsch, “Second-harmonic microscopy of biological tissue,” Opt. Lett., 11 (2), 94 –96 (1986). http://dx.doi.org/10.1364/OL.11.000094 OPLEDP 0146-9592 Google Scholar

6. 

R. LaCombet al., “Phase matching considerations in second harmonic generation from tissues: effects on emission directionality, conversion efficiency and observed morphology,” Opt. Commun., 281 (7), 1823 –1832 (2008). http://dx.doi.org/10.1016/j.optcom.2007.10.040 OPCOB8 0030-4018 Google Scholar

7. 

P. Stolleret al., “Quantitative second-harmonic generation microscopy in collagen,” Appl. Opt., 42 (25), 5209 –5219 (2003). http://dx.doi.org/10.1364/AO.42.005209 APOPAI 0003-6935 Google Scholar

8. 

W. R. ZipfelR. M. WilliamsW. W. Webb, “Nonlinear magic: multiphoton microscopy in the biosciences,” Nat. Biotechnol., 21 (11), 1369 –1377 (2003). http://dx.doi.org/10.1038/nbt899 NABIF9 1087-0156 Google Scholar

9. 

W. R. Zipfelet al., “Live tissue intrinsic emission microscopy using multiphoton-excited native fluorescence and second harmonic generation,” Proc. Natl. Acad. Sci. U. S. A., 100 (12), 7075 –7080 (2003). http://dx.doi.org/10.1073/pnas.0832308100 1091-6490 Google Scholar

10. 

R. M. WilliamsW. R. ZipfelW. W. Webb, “Interpreting second-harmonic generation images of collagen I fibrils,” Biophys. J., 88 (2), 1377 –1386 (2005). http://dx.doi.org/10.1529/biophysj.104.047308 BIOJAU 0006-3495 Google Scholar

11. 

M. HanG. GieseJ. Bille, “Second harmonic generation imaging of collagen fibrils in cornea and sclera,” Opt. Express, 13 (15), 5791 –5797 (2005). http://dx.doi.org/10.1364/OPEX.13.005791 OPEXFF 1094-4087 Google Scholar

12. 

N. Morishigeet al., “Noninvasive corneal stromal collagen imaging using two-photon-generated second-harmonic signals,” J. Cataract. Refract. Surg., 32 (11), 1784 –1791 (2006). http://dx.doi.org/10.1016/j.jcrs.2006.08.027 JCSUEV 0886-3350 Google Scholar

13. 

S.-W. Chuet al., “Thickness dependence of optical second harmonic generation in collagen fibrils,” Opt. Express, 15 (19), 12005 –12010 (2007). http://dx.doi.org/10.1364/OE.15.012005 OPEXFF 1094-4087 Google Scholar

14. 

X. Hanet al., “Second harmonic properties of tumor collagen: determining the structural relationship between reactive stroma and healthy stroma,” Opt. Express, 16 (3), 1846 –1859 (2008). http://dx.doi.org/10.1364/OE.16.001846 OPEXFF 1094-4087 Google Scholar

15. 

Y. Changet al., “Theoretical simulation study of lineraly polarized light on microscopic second-harmonic generation in collagen type I,” J. Biomed. Opt., 14 (4), 044016 (2009). http://dx.doi.org/10.1117/1.3174427 JBOPFO 1083-3668 Google Scholar

16. 

P. StollerK. M. ReiserP. M. Celliers, “Polarization-modulated second harmonic generation in collagen,” Biophys. J., 82 (6), 3330 –3342 (2002). http://dx.doi.org/10.1016/S0006-3495(02)75673-7 BIOJAU 0006-3495 Google Scholar

17. 

A. T. Yehet al., “Selective corneal imaging using combined second-harmonic generation and two-photon excited fluorescence,” Opt. Lett., 27 (23), 2082 –2084 (2002). http://dx.doi.org/10.1364/OL.27.002082 OPLEDP 0146-9592 Google Scholar

18. 

M. Bothet al., “Second harmonic imaging of intrinsic signals in muscle fibers in situ,” J. Biomed. Opt., 9 (5), 882 –892 (2004). http://dx.doi.org/10.1117/1.1783354 JBOPFO 1083-3668 Google Scholar

19. 

F. TiahoG. RecherD. Rouède, “Estimation of helical angles of myosin and collagen by second harmonic generation imaging microscopy,” Opt. Express, 15 (19), 12286 –12295 (2007). http://dx.doi.org/10.1364/OE.15.012286 OPEXFF 1094-4087 Google Scholar

20. 

S. Psilodimitrakopouloset al., “Quantitative discrimination between endogenous SHG sources in mammalian tissue, based on their polarization response,” Opt. Express, 17 (12), 10168 –10176 (2009). http://dx.doi.org/10.1364/OE.17.010168 OPEXFF 1094-4087 Google Scholar

21. 

I. GusachenkoG. LatourM. C. Schanne-Klein, “Polarization-resolved second harmonic microscopy in anisotropic thick tissues,” Opt. Express, 18 (18), 19339 –19352 (2010). http://dx.doi.org/10.1364/OE.18.019339 OPEXFF 1094-4087 Google Scholar

22. 

V. Le Floc’het al., “Monitoring of orientation in molecular ensembles by polarization sensitive nonlinear microscopy,” J. Phys. Chem. B, 107 (45), 12403 –12410 (2003). http://dx.doi.org/10.1021/jp034950t JPCBFK 1520-6106 Google Scholar

23. 

R. M. Plociniket al., “Modular ellipsometric approach for mining structural information from nonlinear optical polarization analysis,” Phys. Rev. B, 72 (12), 125409 (2005). http://dx.doi.org/10.1103/PhysRevB.72.125409 PRBMDO 1098-0121 Google Scholar

24. 

N. J. Begueet al., “Nonlinear optical stokes ellipsometry. 2. experimental demonstration,” J. Phys. Chem. C, 113 (23), 10166 –10175 (2009). http://dx.doi.org/10.1021/jp9013543 1932-7447 Google Scholar

25. 

C.-K. Chouet al., “Polarization ellipticity compensation in polarization second-harmonic generation microscopy without specimen rotation,” J. Biomed. Opt., 13 (1), 014005 (2008). http://dx.doi.org/10.1117/1.2824379 JBOPFO 1083-3668 Google Scholar

26. 

P. Schönet al., “Polarization distortion effects in polarimetric two-photon microscopy,” Opt. Express, 16 (25), 20891 –20901 (2008). http://dx.doi.org/10.1364/OE.16.020891 OPEXFF 1094-4087 Google Scholar

27. 

A. N. Bashkatovet al., “Estimation of wavelength dependence of refractive index of collagen fibers of scleral tissue,” Proc. SPIE, 4162 265 –268 (2000). http://dx.doi.org/10.1117/12.405952 PSISDG 0277-786X Google Scholar

28. 

L. MoreauxO. SandreJ. Mertz, “Membrane imaging by second-harmonic generation microscopy,” J. Opt. Soc. Am. B, 17 (10), 1685 –1694 (2000). http://dx.doi.org/10.1364/JOSAB.17.001685 JOBPDE 0740-3224 Google Scholar

29. 

R. RichardsE. Wolf, “Electromagnetic diffraction in optical systems.2. structure of the image field in an aplanatic system,” Proc. R. Soc. Lond. A. Math. Phys. Sci., 253 (1274), 358 –379 (1959). http://dx.doi.org/10.1098/rspa.1959.0200 PRLAAZ 1364-5021 Google Scholar

30. 

J. M. Bueno, “Polarimetry using liquid-crystal variable retarders: theory and calibration,” J. Opt. A: Pure Appl. Opt., 2 (3), 216 –222 (2000). http://dx.doi.org/10.1088/1464-4258/2/3/308 Google Scholar

31. 

J. MertzL. Moreaux, “Second-harmonic generation by focused excitation of inhomogeneously distributed scatterers,” Opt. Commun., 196 (1–6), 325 –330 (2001). http://dx.doi.org/10.1016/S0030-4018(01)01403-1 OPCOB8 0030-4018 Google Scholar

32. 

E. Hecht, Optics, Higher Education Press, Beijing (2005). Google Scholar

33. 

C. Odinet al., “Orientation fields of nonlinear biological fibrils by second harmonic generation microscopy,” J. Microsc., 229 (1), 32 –38 (2008). http://dx.doi.org/10.1111/jmi.2008.229.issue-1 JMICAR 0022-2720 Google Scholar

34. 

B. VohnsenP. Artal, “Second-harmonic microscopy of ex vivo porcine corneas,” J. Microsc., 232 (1), 158 –163 (2008). http://dx.doi.org/10.1111/jmi.2008.232.issue-1 JMICAR 0022-2720 Google Scholar

35. 

P. Stolleret al., “Quantitative second-harmonic generation microscopy in collagen,” Appl. Opt., 42 (25), 5209 –5219 (2003). http://dx.doi.org/10.1364/AO.42.005209 APOPAI 0003-6935 Google Scholar

36. 

H. Baoet al., “Second harmonic generation imaging via nonlinear endomicroscopy,” Opt. Express, 18 (2), 1255 –1260 (2010). http://dx.doi.org/10.1364/OE.18.001255 OPEXFF 1094-4087 Google Scholar
© 2012 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2012/$25.00 © 2012 SPIE
Oscar del Barco and Juan M. Bueno "Second harmonic generation signal in collagen fibers: role of polarization, numerical aperture, and wavelength," Journal of Biomedical Optics 17(4), 045005 (19 April 2012). https://doi.org/10.1117/1.JBO.17.4.045005
Published: 19 April 2012
Lens.org Logo
CITATIONS
Cited by 17 scholarly publications and 2 patents.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Second-harmonic generation

Polarization

Collagen

Microscopes

Optical fibers

Signal generators

Harmonic generation

Back to Top