This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Measuring the maximally allowed polarization states of the isotropic stochastic gravitational wave background with the ground-based detectors

Hidetoshi Omiya omiya@tap.scphys.kyoto-u.ac.jp Department of Physics, Kyoto University, Kyoto 606-8502, Japan    Naoki Seto seto@tap.scphys.kyoto-u.ac.jp Department of Physics, Kyoto University, Kyoto 606-8502, Japan
Abstract

We discuss the polarizational study of isotropic gravitational wave backgrounds with the second generation detector network, paying special attention to the impacts of adding LIGO-India. The backgrounds can be characterized by at most five spectral components (three parity-even ones and two parity-odd ones). They can be algebraically decomposed through the difference of the corresponding overlap reduction functions defined for the individual spectra. We newly identify two interesting relations between the overlap reduction functions, and these relations generally hamper the algebraic decomposition in the low frequency regime f30f\lesssim 30Hz. We also find that LIGO-India can significantly improve the network sensitives to the odd spectral components.

I Introduction

A stochastic gravitational wave background is one of the primary targets of gravitational wave detectors. There exist a large number of theoretical predictions for generation processes such as an inflationary expansion [1, 2, 3, 4], a phase transition [5, 6], and distant unresolved binaries [7, 8] (for other sources, see [9, 10]). Many of these backgrounds were generated in strong gravity regimes or high energy states and could be a good probe for physics in an extreme environment. Note also that these backgrounds are expected to be highly isotropic.

In General Relativity (GR), we only have the two tensor degrees of freedom, the + and ×\times modes. In contrast, some alternative theories of gravity predict additional polarization modes; the two vector (xx and yy) modes and the two scalar (bb and ll) modes [11]. Therefore, through a polarization study of background, we might detect a signature of modification to GR [12, 13] (see [14, 15, 16, 17, 18, 19] for studies on the polarization of gravitational waves from compact binary). Furthermore, even if GR is not modified at present, a parity violation process in the early universe could generate an asymmetry between right- and left-handed polarization patterns [20, 21, 22, 23, 24, 25, 26, 27].

The cross-correlation analysis is an efficient method for detecting a weak gravitational wave background [28, 29, 30]. By taking products of data streams of noise-independent detector pairs, we can gradually improve the sensitivity to a background by increasing observational time. When the gravitational wave frequency is much longer than the arm lengths of detectors [31], the two scalar modes are observationally non-separable, and we can generally measure the five background spectra IT,IV,IS,WT,I_{T},I_{V},I_{S},W_{T}, and WVW_{V}. The three spectra IT,IV,I_{T},I_{V}, and ISI_{S} represent the total intensity of the tensor, vector, and scalar modes. The remaining two spectra WTW_{T} and WVW_{V} correspond to the Stokes “V” parameters which probe the degrees of circular polarization of the tensor and vector modes. In this paper, we utilize WW for the “V” parameter to avoid confusion with the vector modes. Since the spectra IT,IV,I_{T},I_{V}, and ISI_{S} transform as parity even quantities and the WTW_{T} and WVW_{V} transform as parity odd quantities, we refer to the former three as parity even spectra and the later two as parity odd spectra.

At the correlation analysis, we can measure linear combinations of the five spectra with the five coefficients known as the overlap reduction functions (ORFs). The ORFs characterize the sensitivities to the corresponding spectra and depend on gravitational wave frequency as well as the relative configuration of the two pairwise detectors. We apply the parity even/odd classification also to the five ORFs.

For probing the existence of the anomalous polarization spectra IVI_{V}, ISI_{S}, WTW_{T}, and WVW_{V}, we desire to clean the contribution from the standard spectrum ITI_{T} (see also [32] for a maximum likelihood analysis). In addition, we prefer to break down the four anomalous modes and measure them separately. Our strategy in this paper is to utilize the difference between the five ORFs and algebraically decompose the five spectra by taking appropriate linear combinations of the correlation products from multiple pairs (originally proposed in [33]). We mainly study the prospects of this algebraic scheme with the second generation detector network. We pay special attention to the impacts of adding LIGO-India as the fifth detector.

In the middle of our study, we newly identify two degenerate relations between the ORFs. The first one is for the three even ORFs, and the second one is for the two odd ORFs. These two relations generally limit the performance of the algebraic decomposition in the low frequency regime f30f\lesssim 30Hz. On the other hand, LIGO-India can largely mitigate the damage associated with the degeneracy for the even ORFs, because of its relatively remote location from the two other LIGO detectors. Furthermore, LIGO-India can significantly improve the sensitivities to the odd spectra.

This paper is organized as follows. In Sec. II, we review the polarization decomposition of an isotropic background and present the analytical expressions for the associated ORFs. We also explain our two new findings with respect to the ORFs. In Sec. III, we concretely study the geometry of the second generation terrestrial detector network, including LIGO-India. In Sec. IV, we review the correlation analysis, primarily focusing on the evaluation of the signal-to-noise ratio. In Sec. V, we explain the algebraic decomposition scheme for multiple spectral components. In Secs. VI and  VII, we apply this scheme to the second generation ground-based detector network. We discuss how the sensitivity depends on the target polarization spectra and the network combinations. Finally, in Sec. VIII, we summarize our paper.

II Basic Quantities

Following our preceding work [31] on formal aspects, we first review the basic ingredients for the correlated signals of stochastic backgrounds with ground-based detectors. Since our universe is highly isotropic and homogeneous, the monopole components of the backgrounds are assumed to be our primary target. In addition, because the observed speed of gravitational wave vgv_{g} is close to the speed of light cc, we set vg=cv_{g}=c.

In Sec. II.1, we describe the polarization states of the isotropic backgrounds and introduce the five relevant spectra IT,IV,IS,WT,I_{T},I_{V},I_{S},W_{T}, and WVW_{V}. In Sec. II.2, we discuss the ORFs which characterize the correlated response of pairwise detectors to the backgrounds. In Sec. II.3, we give analytic expressions of the ORFs for the ground-based detectors. In Secs. II.4 and II.5, we discuss their asymptotic behaviors. In Secs. II.6 and II.7, we report our two new findings on the ORFs.

II.1 Polarization states of a stochastic gravitational wave background

We start with the plane wave decomposition of the metric perturbation hijh_{ij} generated by the gravitational waves

hij(t,𝒙)=P𝑑f𝑑𝛀×h~P(f,𝛀)𝒆P,ij(𝛀)e2πif(t𝛀𝒙/c),\displaystyle\begin{aligned} h_{ij}(t,\bm{x})=&\sum_{P}\int df\int d\bm{\Omega}\\ &\times\tilde{h}_{P}(f,\bm{\Omega})\bm{e}_{P,ij}(\bm{\Omega})e^{-2\pi if(t-\bm{\Omega}\cdot\bm{x}/c)}~{},\end{aligned} (1)

where 𝛀\bm{\Omega} is the unit vector for the propagation direction, normalized by 𝑑𝛀=4π\int d\bm{\Omega}=4\pi. Here, 𝒆P\bm{e}_{P} (P=+,×,x,y,bP=+,\times,x,y,b and ll) represent the polarization tensors given by

𝒆+=𝒎𝒎𝒏𝒏,𝒆×=𝒎𝒏+𝒏𝒎,𝒆x=𝛀𝒎+𝒎𝛀,𝒆y=𝛀𝒏+𝒏𝛀,𝒆b=3(𝒎𝒎+𝒏𝒏),𝒆l=3(𝛀𝛀)\displaystyle\begin{aligned} \bm{e}_{+}&=\bm{m}\otimes\bm{m}-\bm{n}\otimes\bm{n}~{},&\bm{e}_{\times}&=\bm{m}\otimes\bm{n}+\bm{n}\otimes\bm{m}~{},\\ \bm{e}_{x}&=\bm{\Omega}\otimes\bm{m}+\bm{m}\otimes\bm{\Omega}~{},&\bm{e}_{y}&=\bm{\Omega}\otimes\bm{n}+\bm{n}\otimes\bm{\Omega}~{},\\ \bm{e}_{b}&=\sqrt{3}(\bm{m}\otimes\bm{m}+\bm{n}\otimes\bm{n})~{},&\bm{e}_{l}&=\sqrt{3}(\bm{\Omega}\otimes\bm{\Omega})\end{aligned} (2)

with the orthonormal vectors 𝒎\bm{m} and 𝒏\bm{n} in addition to 𝛀\bm{\Omega} (see [34] for geometrical interpretation of these modes). Note that our definitions for ebe_{b} and ele_{l} are different from the conventional one such as used in [13] (see also Appendix in [35]). They are written by the standard polar coordinates (θ,ϕ)(\theta,\phi) as

𝛀\displaystyle\bm{\Omega} =(sinθcosϕsinθsinϕcosθ),\displaystyle=\left(\begin{array}[]{c}\sin\theta\cos\phi\\ \sin\theta\sin\phi\\ \cos\theta\end{array}\right)~{}, (6)
𝒎\displaystyle\bm{m} =(cosθcosϕcosθsinϕsinθ),\displaystyle=\left(\begin{array}[]{c}\cos\theta\cos\phi\\ \cos\theta\sin\phi\\ -\sin\theta\end{array}\right)~{}, (10)
𝒏\displaystyle\bm{n} =(sinϕcosϕ0).\displaystyle=\left(\begin{array}[]{c}-\sin\phi\\ \cos\phi\\ 0\end{array}\right)~{}. (14)

In Eq. (2), the labels P=+,×P=+,\times correspond to the tensor (TT) modes, P=x,yP=x,y to the vector (VV) modes, and P=b,lP=b,l to the scalar (SS) modes. Note that GR predicts only the tensor modes. However, numerous alternative theories of gravity allow the existence of the remaining VV and SS modes.

For a stochastic background, the expansion coefficients h~P\tilde{h}_{P} can be regarded as random quantities. Their statistical properties are specified by the power spectrum matrix h~P(f,𝛀)h~P(f,𝛀)\braket{\tilde{h}_{P}(f,\bm{\Omega})\tilde{h}_{P^{\prime}}^{*}(f^{\prime},\bm{\Omega})} with no correlation between T,VT,V and SS modes for statistically isotropic backgrounds [31]. In the case of the tensor modes (P,P=+,×P,P^{\prime}=+,\times), the matrix can be written in terms of the Stokes parameters as [36]

h~P(f,𝛀)h~P(f,𝛀)=12δΩΩδ(ff)×(IT+QTUTiWTUT+iWTITQT)PP.\displaystyle\begin{aligned} \braket{\tilde{h}_{P}(f,\bm{\Omega})\tilde{h}_{P^{\prime}}^{*}(f^{\prime},\bm{\Omega^{\prime}})}=&\frac{1}{2}\delta_{\Omega\Omega^{\prime}}\delta(f-f^{\prime})\\ &\times\left(\begin{array}[]{cc}I_{T}+Q_{T}&U_{T}-iW_{T}\\ U_{T}+iW_{T}&I_{T}-Q_{T}\end{array}\right)_{PP^{\prime}}~{}.\end{aligned} (15)

In the standard literature of polarization, the chiral asymmetry is usually denoted as the Stokes “VV” parameter. In this paper, we apply the notation VV to represent the vector modes, and use WW for the chiral asymmetry. Note that the combinations QT±iUTQ_{T}\pm iU_{T} do not have isotropic components, as understood from their transformation properties [36, 31]. We thus drop them hereafter.

In Eq. (15), we use the coefficients h~P(f,𝛀)\tilde{h}_{P}(f,\bm{\Omega}) for the linear polarization bases (𝒆+,𝒆×)(\bm{e}_{+},\bm{e}_{\times}). However, the physical meaning of the WW parameter becomes transparent by introducing the circular (right- and left-handed) polarization bases given by

𝒆RT\displaystyle\bm{e}_{R}^{T} =12(𝒆++i𝒆×),\displaystyle=\frac{1}{\sqrt{2}}\left(\bm{e}_{+}+i\bm{e}_{\times}\right)~{}, 𝒆LT\displaystyle\bm{e}_{L}^{T} =12(𝒆+i𝒆×),\displaystyle=\frac{1}{\sqrt{2}}\left(\bm{e}_{+}-i\bm{e}_{\times}\right)~{}, (16)

with the corresponding coefficients

h~RT(f,𝛀)\displaystyle\tilde{h}^{T}_{R}(f,\bm{\Omega}) =12(h~+(f,𝛀)ih~×(f,𝛀)),\displaystyle=\frac{1}{\sqrt{2}}\left(\tilde{h}_{+}(f,\bm{\Omega})-i\tilde{h}_{\times}(f,\bm{\Omega})\right)~{}, (17)
h~LT(f,𝛀)\displaystyle\tilde{h}^{T}_{L}(f,\bm{\Omega}) =12(h~+(f,𝛀)+ih~×(f,𝛀)).\displaystyle=\frac{1}{\sqrt{2}}\left(\tilde{h}_{+}(f,\bm{\Omega})+i\tilde{h}_{\times}(f,\bm{\Omega})\right). (18)

We then have

IT\displaystyle I_{T} =h~RTh~RT+h~LTh~LT,\displaystyle=\braket{\tilde{h}^{T}_{R}\tilde{h}_{R}^{T*}}+\braket{\tilde{h}^{T}_{L}\tilde{h}_{L}^{T*}}~{}, (19)
WT\displaystyle W_{T} =h~RTh~RTh~LTh~LT,\displaystyle=\braket{\tilde{h}^{T}_{R}\tilde{h}_{R}^{T*}}-\braket{\tilde{h}^{T}_{L}\tilde{h}_{L}^{T*}}, (20)

omitting apparent delta functions. These expressions show that the spectra ITI_{T} and WTW_{T} characterize the total and asymmetry of the amplitudes of the right- and left-handed polarization patterns of the tensor modes. Since the parity transformation interchanges the right-and left-handed waves, we resultantly have IT=ITI_{T}^{\prime}=I_{T} and WT=WTW_{T}^{\prime}=-W_{T} ( representing parity transformed quantities).

For the vector modes, we can repeat almost the same arguments as Eqs. (15)-(18) and obtain

IV\displaystyle I_{V} =h~RVh~RV+h~LVh~LV,\displaystyle=\braket{\tilde{h}^{V}_{R}\tilde{h}_{R}^{V*}}+\braket{\tilde{h}^{V}_{L}\tilde{h}_{L}^{V*}}~{}, (21)
WV\displaystyle W_{V} =h~RVh~RVh~LVh~LV\displaystyle=\braket{\tilde{h}^{V}_{R}\tilde{h}_{R}^{V*}}-\braket{\tilde{h}^{V}_{L}\tilde{h}_{L}^{V*}} (22)

with the correspondences IV=IVI_{V}^{\prime}=I_{V} and WV=WVW_{V}^{\prime}=-W_{V} for the parity transformation.

For the scalar modes (P,P=b,lP,P^{\prime}=b,l), considering their potential correlation, we can generally put

h~P(f,𝛀)h~P(f,𝛀)=12δΩΩδ(ff)×(IbCSCSIl)PP.\displaystyle\begin{aligned} \braket{\tilde{h}_{P}(f,\bm{\Omega})\tilde{h}_{P^{\prime}}^{*}(f^{\prime},\bm{\Omega^{\prime}})}=&\frac{1}{2}\delta_{\Omega\Omega^{\prime}}\delta(f-f^{\prime})\\ &\times\left(\begin{array}[]{cc}I_{b}&C_{S}\\ C_{S}^{*}&I_{l}\end{array}\right)_{PP^{\prime}}~{}.\end{aligned} (23)

described by the four real parameters in the power spectra. In reality, as long as the low frequency approximation is valid (f(2πL/c)1f\ll(2\pi L/c)^{-1}, LL: the arm length), only the combination

IS12(Ib+IlCSCS),\displaystyle I_{S}\equiv\frac{1}{2}(I_{b}+I_{l}-C_{S}-C_{S}^{*})~{}, (24)

appears in the correlation analysis [31]. Therefore, in the following, we keep only ISI_{S} for the scalar modes. Because of its spin-0 nature, we also have IS=ISI_{S}^{\prime}=I_{S} for the parity transformation.

Up to now, we see that an isotropic background is characterized by the five quantities IT,IV,IS,WT,I_{T},I_{V},I_{S},W_{T}, and WVW_{V}. Here, we introduce another commonly used representation for the magnitudes of these spectra. In GR, the amplitude IT(f)I_{T}(f) can be simply related to the energy density of the background. More specifically, with the Hubble parameter H0H_{0}, we have the relation

ΩGWIT(f)=(32π33H02)f3IT(f)\displaystyle\Omega_{GW}^{I_{T}}(f)=\left(\frac{32\pi^{3}}{3H_{0}^{2}}\right)f^{3}I_{T}(f)~{} (25)

for the energy density of the background per logarithmic frequency (normalized by critical density of universe) [28, 29]. In a modified theory of gravity, the relation (25) for the energy density might be invalid [37]. However, we are not directly interested in the actual energy density of the backgrounds, and thus continue to use Eq. (25) as the definition of ΩGWIT(f)\Omega_{GW}^{I_{T}}(f). Similarly, we use the effective energy densities

ΩGWIV(f)\displaystyle\Omega_{GW}^{I_{V}}(f) (32π33H02)f3IV(f),\displaystyle\equiv\left(\frac{32\pi^{3}}{3H_{0}^{2}}\right)f^{3}I_{V}(f)~{}, (26)
ΩGWIS(f)\displaystyle\Omega_{GW}^{I_{S}}(f) (32π33H02)f3IS(f),\displaystyle\equiv\left(\frac{32\pi^{3}}{3H_{0}^{2}}\right)f^{3}I_{S}(f)~{}, (27)
ΩGWWT(f)\displaystyle\Omega_{GW}^{W_{T}}(f) (32π33H02)f3WT(f),\displaystyle\equiv\left(\frac{32\pi^{3}}{3H_{0}^{2}}\right)f^{3}W_{T}(f)~{}, (28)
ΩGWWV(f)\displaystyle\Omega_{GW}^{W_{V}}(f) (32π33H02)f3WV(f).\displaystyle\equiv\left(\frac{32\pi^{3}}{3H_{0}^{2}}\right)f^{3}W_{V}(f)~{}. (29)

If the left handed modes dominate the right handed ones, we have ΩGWWT(f)<0\Omega_{GW}^{W_{T}}(f)<0 (and ΩGWWV(f)<0\Omega_{GW}^{W_{V}}(f)<0).

II.2 Correlation Analysis

Now we discuss how to detect the five spectral components by using multiple interferometers. In the low frequency regime (f(2πL/c)1f\ll(2\pi L/c)^{-1}), the response of an interferometer AA (at the position 𝒙A\bm{x}_{A}) can be modeled as [31]

hA(f)=𝑫Aijh~ij(f,𝒙A),\displaystyle h_{A}(f)=\bm{D}_{A}^{ij}\tilde{h}_{ij}(f,\bm{x}_{A})~{}, (30)

with the beam pattern function

𝑫A=𝒖A𝒖A𝒗A𝒗A2.\displaystyle\bm{D}_{A}=\frac{\bm{u}_{A}\otimes\bm{u}_{A}-\bm{v}_{A}\otimes\bm{v}_{A}}{2}~{}. (31)

Here, h~ij(f,𝒙A)\tilde{h}_{ij}(f,\bm{x}_{A}) is the metric perturbation of the background at the detector, and the two unit vectors 𝒖A\bm{u}_{A} and 𝒗A\bm{v}_{A} represent the two arm directions of the detector.

By correlating data streams of multiple detectors, we can statistically amplify the background signals relative to the detector noises (closely discussed in Sec. IV). We denote the correlation product of two detector AA and BB by

CAB(f)hA(f)hB(f)\displaystyle C_{AB}(f)\equiv\braket{h_{A}(f)h_{B}^{*}(f)}~{} (32)

(again omitting delta functions). Leaving only the monopole components of the background, we obtain

CAB(f)=4π5(P=T,V,SγIPIP+P=T,VγWPWP).\displaystyle\begin{aligned} C_{AB}(f)&=\frac{4\pi}{5}\left(\sum_{P=T,V,S}\gamma^{I_{P}}I_{P}+\sum_{P=T,V}\gamma^{W_{P}}W_{P}\right)~{}.\end{aligned} (33)

Here, γIP\gamma^{I_{P}} and γWP\gamma^{W_{P}} are the ORFs which characterize the correlated response of two detectors to the relevant components of an isotropic background. They are written as

γABIP\displaystyle\gamma^{I_{P}}_{AB} (f)𝑫A,ij𝑫B,klΓijklIP,\displaystyle(f)\equiv\bm{D}_{A,ij}\bm{D}_{B,kl}\Gamma^{I_{P}}_{ijkl}~{}, (34)
γABWP\displaystyle\gamma^{W_{P}}_{AB} (f)𝑫A,ij𝑫B,klΓijklWP,\displaystyle(f)\equiv\bm{D}_{A,ij}\bm{D}_{B,kl}\Gamma^{W_{P}}_{ijkl}~{}, (35)

with the angular integrals

ΓijklIT\displaystyle\Gamma^{I_{T}}_{ijkl} =58π𝑑𝛀(e+,ije+,kl+e×,ije×,kl)eiy𝛀𝒅^,\displaystyle=\frac{5}{8\pi}\int d\bm{\Omega}(e_{+,ij}e_{+,kl}+e_{\times,ij}e_{\times,kl})e^{iy\bm{\Omega}\cdot\hat{\bm{d}}}~{}, (36)
ΓijklIV\displaystyle\Gamma^{I_{V}}_{ijkl} =58π𝑑𝛀(ex,ijex,kl+ey,ijey,kl)eiy𝛀𝒅^,\displaystyle=\frac{5}{8\pi}\int d\bm{\Omega}(e_{x,ij}e_{x,kl}+e_{y,ij}e_{y,kl})e^{iy\bm{\Omega}\cdot\hat{\bm{d}}}~{}, (37)
ΓijklIS\displaystyle\Gamma^{I_{S}}_{ijkl} =58π𝑑𝛀(eb,ijeb,kl+el,ijel,kl)eiy𝛀𝒅^,\displaystyle=\frac{5}{8\pi}\int d\bm{\Omega}(e_{b,ij}e_{b,kl}+e_{l,ij}e_{l,kl})e^{iy\bm{\Omega}\cdot\hat{\bm{d}}}~{}, (38)
ΓijklWT\displaystyle\Gamma^{W_{T}}_{ijkl} =5i8π𝑑𝛀(e+,ije×,kle×,ije+,kl)eiy𝛀𝒅^,\displaystyle=-\frac{5i}{8\pi}\int d\bm{\Omega}(e_{+,ij}e_{\times,kl}-e_{\times,ij}e_{+,kl})e^{iy\bm{\Omega}\cdot\hat{\bm{d}}}~{}, (39)
ΓijklWV\displaystyle\Gamma^{W_{V}}_{ijkl} =5i8π𝑑𝛀(ex,ijey,kley,ijex,kl)eiy𝛀𝒅^.\displaystyle=-\frac{5i}{8\pi}\int d\bm{\Omega}(e_{x,ij}e_{y,kl}-e_{y,ij}e_{x,kl})e^{iy\bm{\Omega}\cdot\hat{\bm{d}}}~{}. (40)

Here, we put d|𝒙A𝒙B|d\equiv|{\bm{x}}_{A}-{\bm{x}}_{B}|, 𝒅^(𝒙A𝒙B)/d\hat{\bm{d}}\equiv({\bm{x}}_{A}-{\bm{x}}_{B})/d and y=2πfd/cy=2\pi fd/c.

As already in Eq. (33), we will use the label PP for the polarization modes P=(T,V,S)P=(T,V,S), extending it from the original patterns P=(+,×,x,y,b,l)P=(+,\times,x,y,b,l). In addition, we introduce the label QQ to represent all the five spectral modes (IT,IV,IS,WT,WV)(I_{T},I_{V},I_{S},W_{T},W_{V}) of interest. For notational simplicity, we also omit the labels for the detectors in obvious cases.

II.3 ORFs for ground-based detectors

Refer to caption
Figure 1: The relative geometry of the ground-based detector pair AA and BB. The two detectors are on the same great circle and their detector planes are tangential to the earth sphere. The opening angle β\beta is measured from the center of the Earth. The angles σA\sigma_{A} and σB\sigma_{B} correspond to the orientations of the bisectors of the two arms (dotted line) measured counter clock wisely relative to the great circle.

Now we focus on the ground-based detectors that are assumed to be tangential to the Earth sphere of the radius RE=6400R_{E}=6400km. As shown in Fig. 1, the relative geometry of two interferometers AA and BB are fully characterized by three angles β\beta, σA\sigma_{A} and σB\sigma_{B} (following the convention in [28]). The angle β\beta represents the opening angle between the two detectors, measured from the center of the Earth, and we have d=2REsin(β/2)d=2R_{E}\sin(\beta/2). Meanwhile, the angle σA\sigma_{A} shows the orientation of the bisector of the two arms of the detector AA measured counterclockwise relative to the great circle joining the two detectors. The angle σB\sigma_{B} is defined similarly. Below, instead of σA\sigma_{A} and σB\sigma_{B}, we use the angles Δ\Delta and δ\delta

Δ\displaystyle\Delta σA+σB2,\displaystyle\equiv\frac{\sigma_{A}+\sigma_{B}}{2}~{}, δ\displaystyle\delta σAσB2,\displaystyle\equiv\frac{\sigma_{A}-\sigma_{B}}{2}~{}, (41)

following the standard convention.

The close expressions of the ORFs are presented in [31] as

γIP\displaystyle\gamma^{I_{P}} =ΘΔP(y,β)cos4Δ+ΘδP(y,β)cos4δ,\displaystyle=\Theta_{\Delta}^{P}(y,\beta)\cos 4\Delta+\Theta_{\delta}^{P}(y,\beta)\cos 4\delta~{}, (P\displaystyle(P =T,V,S),\displaystyle=T,V,S)~{}, (42)
γWP\displaystyle\gamma^{W_{P}} =ΞP(y,β)sin4Δ,\displaystyle=\Xi^{P}(y,\beta)\sin 4\Delta~{}, (P\displaystyle(P =T,V).\displaystyle=T,V)~{}. (43)

Here the angles δ\delta and Δ\Delta appear only in the forms cos4δ,cos4Δ,\cos 4\delta,\cos 4\Delta, and sin4Δ\sin 4\Delta, reflecting certain symmetries [38]. The coefficients ΞP,ΘΔP,\Xi^{P},\Theta_{\Delta}^{P}, and ΘδP\Theta_{\delta}^{P} are given by

ΘΔT(y,β)\displaystyle\Theta^{T}_{\Delta}(y,\beta) =sin4(β2)j0(y)\displaystyle=-\sin^{4}\left(\frac{\beta}{2}\right)j_{0}(y) (44)
556(9+8cosβ+cos2β)j2(y)\displaystyle-\frac{5}{56}(-9+8\cos\beta+\cos 2\beta)j_{2}(y) (45)
1896(169+108cosβ+3cos2β)j4(y),\displaystyle-\frac{1}{896}(169+108\cos\beta+3\cos 2\beta)j_{4}(y)~{}, (46)
ΘΔV(y,β)\displaystyle\Theta^{V}_{\Delta}(y,\beta) =sin4(β2)j0(y)\displaystyle=-\sin^{4}\left(\frac{\beta}{2}\right)j_{0}(y) (47)
+5112(9+8cosβ+cos2β)j2(y)\displaystyle+\frac{5}{112}(-9+8\cos\beta+\cos 2\beta)j_{2}(y) (48)
+1224(169+108cosβ+3cos2β)j4(y),\displaystyle+\frac{1}{224}(169+108\cos\beta+3\cos 2\beta)j_{4}(y)~{}, (49)
ΘΔS(y,β)\displaystyle\Theta^{S}_{\Delta}(y,\beta) =sin4(β2)j0(y)\displaystyle=-\sin^{4}\left(\frac{\beta}{2}\right)j_{0}(y) (50)
+556(9+8cosβ+cos2β)j2(y)\displaystyle+\frac{5}{56}(-9+8\cos\beta+\cos 2\beta)j_{2}(y) (51)
3448(169+108cosβ+3cos2β)j4(y),\displaystyle-\frac{3}{448}(169+108\cos\beta+3\cos 2\beta)j_{4}(y)~{}, (52)
ΘδT(y,β)\displaystyle\Theta^{T}_{\delta}(y,\beta) =cos4(β2)(j0(y)+57j2(y)+3112j4(y)),\displaystyle=\cos^{4}\left(\frac{\beta}{2}\right)\left(j_{0}(y)+\frac{5}{7}j_{2}(y)+\frac{3}{112}j_{4}(y)\right)~{}, (53)
ΘδV(y,β)\displaystyle\Theta^{V}_{\delta}(y,\beta) =cos4(β2)(j0(y)514j2(y)328j4(y)),\displaystyle=\cos^{4}\left(\frac{\beta}{2}\right)\left(j_{0}(y)-\frac{5}{14}j_{2}(y)-\frac{3}{28}j_{4}(y)\right)~{}, (54)
ΘδS(y,β)\displaystyle\Theta^{S}_{\delta}(y,\beta) =cos4(β2)(j0(y)57j2(y)+956j4(y)),\displaystyle=\cos^{4}\left(\frac{\beta}{2}\right)\left(j_{0}(y)-\frac{5}{7}j_{2}(y)+\frac{9}{56}j_{4}(y)\right)~{}, (55)
ΞT(y,β)\displaystyle\Xi^{T}(y,\beta) =sin(β2)((1cosβ)j1(y)7+3cosβ8j3(y)),\displaystyle=\sin\left(\frac{\beta}{2}\right)\left((1-\cos\beta)j_{1}(y)-\frac{7+3\cos\beta}{8}j_{3}(y)\right)~{}, (56)
ΞV(y,β)\displaystyle\Xi^{V}(y,\beta) =12sin(β2)((1cosβ)j1(y)+7+3cosβ2j3(y))\displaystyle=\frac{1}{2}\sin\left(\frac{\beta}{2}\right)\left((1-\cos\beta)j_{1}(y)+\frac{7+3\cos\beta}{2}j_{3}(y)\right) (57)

with the spherical Bessel functions jn(y)j_{n}(y).

II.4 Asymptotic Behaviors at yy\to\infty

In this subsection, we briefly discuss the asymptotic profiles of the ORFs at yy\to\infty, based on Eqs. (42)-(57).

For the spherical Bessel functions, at large yy, we have the following correspondences

j2l(y)\displaystyle j_{2l}(y) sinyy,\displaystyle\propto\frac{\sin y}{y}~{}, j2l+1(y)\displaystyle j_{2l+1}(y) cosyy,\displaystyle\propto\frac{\cos y}{y}~{}, (58)

Then, we can put

γIP\displaystyle\gamma^{I_{P}} CIPsinyy,\displaystyle\to C_{I_{P}}\frac{\sin y}{y}~{}, γWP\displaystyle\gamma^{W_{P}} CWPcosyy\displaystyle\to C_{W_{P}}\frac{\cos y}{y} (59)

with the coefficients CIPC_{I_{P}} and CWPC_{W_{P}} presented shortly. Roughly speaking, these relations show the phase offset of π/2\sim\pi/2 (as in the combination of siny\sin y and cosy\cos y), depending on the two parity types of the background spectra IPI^{P} and WPW^{P}.

We can readily evaluate the coefficients CIPC_{I_{P}} and CWPC_{W_{P}} as follows;

CIT\displaystyle C_{I_{T}} =5128(8cos4(β2)cos4δ\displaystyle=\frac{5}{128}\Bigl{(}8\cos^{4}\left(\frac{\beta}{2}\right)\cos 4\delta (60)
(cos2β28cosβ+35)cos4Δ),\displaystyle-(\cos 2\beta-28\cos\beta+35)\cos 4\Delta\Bigr{)}~{}, (61)
CIV\displaystyle C_{I_{V}} =58cos2(β2)(2cos2(β2)cos4δ\displaystyle=\frac{5}{8}\cos^{2}\left(\frac{\beta}{2}\right)\Bigl{(}2\cos^{2}\left(\frac{\beta}{2}\right)\cos 4\delta (62)
(cosβ3)cos4Δ),\displaystyle-(\cos\beta-3)\cos 4\Delta\Bigr{)}~{}, (63)
CIS\displaystyle C_{I_{S}} =158cos4(β2)(cos4δcos4Δ),\displaystyle=\frac{15}{8}\cos^{4}\left(\frac{\beta}{2}\right)(\cos 4\delta-\cos 4\Delta)~{}, (64)
CWT\displaystyle C_{W_{T}} =516(sin(3β2)+7sin(β2))sin4Δ,\displaystyle=-\frac{5}{16}\left(-\sin\left(\frac{3\beta}{2}\right)+7\sin\left(\frac{\beta}{2}\right)\right)\sin 4\Delta~{}, (65)
CWV\displaystyle C_{W_{V}} =52sin(β2)cos2(β2)sin4Δ.\displaystyle=\frac{5}{2}\sin\left(\frac{\beta}{2}\right)\cos^{2}\left(\frac{\beta}{2}\right)\sin 4\Delta~{}. (66)

We have CWTCWV0C_{W_{T}}\cdot C_{W_{V}}\leq 0. Notice that CWTC_{W_{T}} and CWVC_{W_{V}} vanish at β=0\beta=0. Two detectors on a plane are apparently mirror symmetric, and thus blind to the parity odd polarizations.

II.5 Asymptotic Behaviors at y0y\to 0

At the opposite limit, y0y\to 0, we have

γIT,V,S(y)\displaystyle\gamma^{I_{T,V,S}}(y) 2DA,ijDBij\displaystyle\to 2D_{A,ij}D_{B}^{ij} (67)
=sin4(β2)cos4Δ+cos4(β2)cos4δ,\displaystyle=-\sin^{4}\left(\frac{\beta}{2}\right)\cos 4\Delta+\cos^{4}\left(\frac{\beta}{2}\right)\cos 4\delta~{}, (68)
γWT(y)\displaystyle\gamma^{W_{T}}(y) 2sin3(β2)ysin4Δ,\displaystyle\to 2\sin^{3}\left(\frac{\beta}{2}\right)y\sin 4\Delta~{}, (69)
γWV(y)\displaystyle\gamma^{W_{V}}(y) sin3(β2)ysin4Δ.\displaystyle\to\sin^{3}\left(\frac{\beta}{2}\right)y\sin 4\Delta~{}. (70)

The first expression shows the degeneracy of the parity even ORFs. Meanwhile, the parity odd ORFs vanish at y0y\to 0, due to the parity symmetry of a network at the same place with d=0d=0 [31]. Thus, a network becomes blind to the parity odd polarizations for small yy.

II.6 Trinity degeneracy of even ORFs at the sub-leading order O(y2)O(y^{2})

At the sub-leading order O(y2)O(y^{2}) (or equivalently O(f2)O(f^{2})), we can easily confirm a cancellation for the three even ORFs and have

γIT(y)4γIV(y)+3γIS(y)=O(y4).\gamma^{I_{T}}(y)-4\gamma^{I_{V}}(y)+3\gamma^{I_{S}}(y)=O(y^{4})~{}. (71)

This trinity degeneracy will later play an important role in the spectral decomposition of the three even spectra.

II.7 Degeneracy of odd ORFs at 13Hz

For detectors on the Earth, we can put y=ζsin(β/2)y=\zeta\sin(\beta/2) with ζ4πREf/c\zeta\equiv 4\pi R_{E}f/c. In Fig. 2, we present a contour plot for the ratio between the odd ORFs

γWTγWV=ΞT(y,β)ΞV(y,β)Θ(ζ,β).\frac{\gamma_{W_{T}}}{\gamma_{W_{V}}}=\frac{\Xi^{T}(y,\beta)}{\Xi^{V}(y,\beta)}\equiv\Theta(\zeta,\beta). (72)

At the left end, we can see the limit limζ0Θ(ζ,β)=2\lim_{\zeta\to 0}\Theta(\zeta,\beta)=2 following from Eqs. (69) and (70).

Surprisingly, the function Θ\Theta depends very weakly on β\beta around ζ=3.57\zeta=3.57, as shown with the almost vertical contour Θ=1.26\Theta=1.26 in Fig. 2. Indeed, along this contour, the variation of ζ\zeta is within ±0.01\pm 0.01. Later, we will find that the odd spectral decomposition practically collapses around ζ=3.57\zeta=3.57, corresponding to f=13f=13Hz for the Earth’s radius RE=6400R_{E}=6400km. This anathematic frequency is intrinsic to ground-based detectors.

In space, we might realize a detector network composed by multiple LISA-like units orbiting around the Sun [39, 40] (see also [41]). For their typical orbital configuration, we need at least three separated units for fully decomposing the five polarization spectra, and these units contact with a virtual sphere of radius 1.15 a.u. [42, 35, 43] (see also [44]). In this case, the anathematic frequency becomes 0.57mHz.

Refer to caption
Figure 2: The contour plot for the ratio Θ(ζ,β)\Theta(\zeta,\beta). We have the limit limζ0Θ=2\lim_{\zeta\to 0}\Theta=2 and almost vertical contour line around ζ=3.575\zeta=3.575 (corresponding to 13Hz for ground-based detectors).

III second generation detector network

Table 1: The latitudes, longitudes, and orientations of the five ground-based detectors in units of degree. The angle α\alpha is the orientation angle of the bisector of the two arms measured from the local east at each detector222https://git.ligo.org/.
detector latitude longitude α\alpha
KAGRA(K) 36.41 137.31 74.60
LIGO-I(I) 19.61 77.03 162.62
LIGO-H(H) 46.45 -119.41 171.00
LIGO-L(L) 30.56 -90.8 242.17
Virgo(V) 43.63 10.50 115.57
Table 2: (Upper right) The opening angle β\beta (in units of degree) of the detector pairs, measured from the center of the Earth. (Lower left) The values of (cos4δ,cos4Δ,sin4Δ)(\cos 4\delta,\cos 4\Delta,\sin 4\Delta).
KAGRA LIGO-I LIGO-H LIGO-L Virgo
KAGRA * 54.89 72.37 99.27 86.52
LIGO-I (-0.41,0.63,0.78) * 112.28 128.47 59.79
LIGO-H (0.99,-0.34,0.94) (0.75,0.47,-0.88) * 27.22 79.62
LIGO-L (-1.00,0.19,-0.98) (-0.80,-0.06,1.00) (-1.00,-0.40,-0.91) * 76.76
Virgo (-0.60,0.87,0.50) (-0.99,0.14,-0.99) (-0.43,-0.80,-0.60) (-0.31,0.86,-0.50) *
Table 3: The expansion coefficients (CIT,CIV,CIS)(C_{I_{T}},C_{I_{V},}C_{I_{S}}) (upper right) and (CWT,CWV)(C_{W_{T}},C_{W_{V}}) (lower left).
KAGRA LIGO-I LIGO-H LIGO-L Virgo
KAGRA * (-0.54,0.43,-1.22) (0.48,0.15,1.06) (-0.34,-0.06,-0.39) (-1.1,0.64,-0.77)
LIGO-I (-0.54,0.70) * (-0.80,0.40,0.05) (0.11,-0.06,-0.05) (-0.29,-0.53,-1.20)
LIGO-H (-0.93,0.90) (1.55,-0.57) * (-0.11,-1.62,-1.00) (0.86,-1.02,0.24)
LIGO-L (1.48,-0.78) (-2.04,0.42) (0.28,-0.51) * (-0.97,0.77,-0.83)
Virgo (-0.63,0.45) (0.77,-0.93) (0.68,-0.57) (0.54,-0.48) *

From now on, we mainly discuss the ground-based detector networks composed by the following five second generation interferometers; LIGO-Handford (H), LIGO-India (I), KAGRA (K), LIGO-Livingston (L) and Virgo (V). We present their basic angular parameters in Table 1.

From these five interferometers, we can make C25=10\rm{}_{5}C_{2}=10 pairs and introduce the abstract index uu to represent these ten pairs {HI,HK,,LV}\{{\rm HI,HK,\cdots,LV}\}. Their relative geometrical parameters are presented in Table 2.

Since each pair has the five ORFs γuQ(f)\gamma_{u}^{Q}(f) (Q;IT,IV,IS,WT,WVQ;I_{T},I_{V},I_{S},W_{T},W_{V}), the total number of ORFs is 50. In Fig. 3, we present all of them at a clip. Later, we will come to deal with the sums of their products such as uγuQ(f)γuQ(f)\sum_{u}\gamma_{u}^{Q}(f)\gamma_{u}^{Q^{\prime}}(f), and the collective behaviours of these large number of ORFs would be important there.

As explained in Sec. II.E, at f=0f=0, we have the degeneracies γuIT=γuIV=γuIS\gamma_{u}^{I_{T}}=\gamma_{u}^{I_{V}}=\gamma_{u}^{I_{S}} and γuWT=γuWV=0\gamma_{u}^{W_{T}}=\gamma_{u}^{W_{V}}=~{}0. In Fig. 3, we can easily identify the three conspicuous curves starting from γuQ0.9\gamma_{u}^{Q}\simeq-0.9 at f=0f=0. These are the even ORFs of the HL pair. This pair is designed to have a large overlap with cos4δ1\cos 4\delta\simeq-1. In Fig. 4, its five ORFs are presented, showing loose oscillation patterns due to the small separation angle β\beta. The small angle β\beta also suppresses the amplitudes of the odd ORFs, in contrast to the even ones (see Sec. II.4 and II.5).

Meanwhile, the HI and IL pairs have large separation angles β\beta and thus provide relatively large value

yfsin(β/2)y\propto f\sin(\beta/2) (73)

for a given frequency ff. This will help us to use the higher order correction terms of the variables yy (e.g., breaking the spectral degeneracy). Together with the preferred relative orientation |sin4Δ|1|\sin 4\Delta|\sim 1, these pairs also have good sensitivities to the odd parity spectra WTW_{T} and WVW_{V}.

As examples of typical pairs, in Fig. 5, we show the ORFs of the LV-pair. In the bottom panel, we compare the asymptotic profiles discussed in Sec. II.D. At f80f\gtrsim 80Hz (y4πy\gtrsim 4\pi), they show reasonable agreements with the original curves. Accordingly, in the upper panel, we can see the phase offset π/2\sim\pi/2 between the odd and even ORFs there. In Table 3, we present the asymptotic coefficients CQC_{Q} for the ten pairs.

Refer to caption
Refer to caption
Figure 3: All the 50 ORFs formed from the five interfeometers H, I, K, L and V. In the upper panel, the three curves starting from 0.9-0.9 correspond to the HL pair.
Refer to caption
Figure 4: All the five ORFs of the HL pair with y=6.3(f/100Hz)y=6.3(f/100{\rm Hz}). The solid lines correspond to the even ORFs. The dashed lines show the odd ones.
Refer to caption
Refer to caption
Figure 5: The ORFs of the LV pair with y=16.65(f/100Hz)y=16.65(f/100{\rm Hz}). (Top) All the five ORFs. The solid lines correspond the even ORFs. The dashed lines are the odd ones. (Bottom) The ORFs for the two tensor modes ITI_{T} and WTW_{T}. The solid ones are the original expressions, and the dotted lines show their asymptotic profiles Eq. (59) with the coefficients (CIT,CWT)(C_{I_{T}},C_{W_{T}}) given in Table 3.

IV Correlation analysis with ground-based detectors

Up to this point, we only considered the response of detectors to stochastic backgrounds. In reality, the data streams of the detectors are contaminated by the detector noises. As we see below, the correlation analysis is a powerful framework to coherently amplify the background signals relative to the noises [28, 29].

Under the existence of the detector noises, the outputs of two detectors AA and BB can be modeled as

sA(f)\displaystyle s_{A}(f) =hA(f)+nA(f),\displaystyle=h_{A}(f)+n_{A}(f)~{}, sB(f)\displaystyle s_{B}(f) =hB(f)+nB(f).\displaystyle=h_{B}(f)+n_{B}(f)~{}. (74)

Here, hA,Bh_{A,B} are the signals from stochastic backgrounds (see Eq. (30)) and nA,Bn_{A,B} are the detector noises. In this paper, we assume the noises nA,Bn_{A,B} to be stationary, Gaussian, and mutually independent. In addition, the signals are assumed to be much smaller than the noises, namely |hA,B||nA,B||h_{A,B}|\ll|n_{A,B}| (the weak signal condition). Then the covariance of the detector noises is given by

nA(f)nB(f)=δAB2NA(f)δ(ff),\displaystyle\braket{n_{A}(f)n_{B}^{*}(f^{\prime})}=\frac{\delta_{AB}}{2}N_{A}(f)\delta(f-f^{\prime})~{}, (75)

where NAN_{A} is the noise power spectrum.

As a preparation of the correlation analysis, let us take the product of the two outputs of pairwise detectors (u=ABu=AB) as (again omitting the delta functions)

μu(f)Re[sA(f)sB(f)].\displaystyle\mu_{u}(f)\equiv\mathrm{Re}[s_{A}(f)s_{B}^{*}(f)]~{}. (76)

Here we extracted the real part. This is because, we know the following relation

sA(f)sB(f)\displaystyle\braket{s_{A}(f)s^{*}_{B}(f)} =\displaystyle= hA(f)hB(f)+hA(f)nB(f)\displaystyle\braket{h_{A}(f)h_{B}(f)}+\braket{h_{A}(f)n_{B}(f)} (78)
+nA(f)hB(f)+nA(f)nB(f)]\displaystyle+\braket{n_{A}(f)h_{B}(f)}+\braket{n_{A}(f)n_{B}(f)}]
=\displaystyle= hA(f)hB(f)\displaystyle\braket{h_{A}(f)h_{B}(f)} (79)
=\displaystyle= Cu(f)Real\displaystyle C_{u}(f)\in{\rm Real} (80)

for the expectation value (using the statistical independence between hA,Bh_{A,B} and nA,Bn_{A,B}). As we see shortly, this projection can reduce the associated noise level [45].

The variance can be calculated similarly. Under the weak signal condition (|hA(f)||nA(f)||h_{A}(f)|\ll|n_{A}(f)|), we have

𝒩u(f)=\displaystyle\mathcal{N}_{u}(f)= μu2μu2μu2\displaystyle\braket{\mu_{u}^{2}}-\braket{\mu_{u}}^{2}\sim\braket{\mu_{u}^{2}} (81)
=\displaystyle= 14(sAsB+sAsB)(f)(sAsB+sAsB)(f)\displaystyle\frac{1}{4}\braket{(s_{A}s_{B}^{*}+s_{A}^{*}s_{B})(f)(s_{A}s_{B}^{*}+s_{A}^{*}s_{B})(f)} (82)
\displaystyle\sim 12nA(f)nB(f)nA(f)nB(f)\displaystyle\frac{1}{2}\braket{n_{A}(f)n_{B}^{*}(f)n_{A}^{*}(f)n_{B}(f)} (83)
=\displaystyle= 18NA(f)NB(f)\displaystyle\frac{1}{8}N_{A}(f)N_{B}(f)~{} (84)

with 𝒩u(f)μu(f)\sqrt{\mathcal{N}_{u}(f)}\gg\left\langle\mu_{u}(f)\right\rangle. Note that we have the additional factor 212^{-1} due to the real projection (76).

The basic idea of the correlation analysis is to coherently amplify the background signal relative to the noise, by using a large number of Fourier modes, after a long observational time. We now explain this by deriving Eqs. (90) and (91).

To deal with the frequency dependence, we first divide the Fourier modes into NN bins (B1,B2,,Bρ,,BN)(B_{1},B_{2},\cdots,B_{\rho},\cdots,B_{N}) characterized by the central frequencies fρf_{\rho} and a fixed width δf\delta f [45]. We take δf\delta f to be much smaller than fρf_{\rho}, such that involved quantities (e.g. IP(f),WP(f),I^{P}(f),W^{P}(f), and γIP,WP(f)\gamma^{I_{P},W_{P}}(f)) are nearly the same in each bin. Meanwhile, we also set the width δf\delta f to be much larger than the frequency resolution Tobs1T_{\rm obs}^{-1} determined by the observation time TobsT_{\rm obs} (i.e. the number of the modes Tobsδf1T_{\rm obs}\delta f\gg 1 in each bin).

Now, we sum up the product μu\mu_{u} in each bin as

μuρ=\displaystyle\mu_{u}^{\rho}= fBρRe[sA(f)sB(f)]\displaystyle\sum_{f\in B_{\rho}}\mathrm{Re}[s_{A}(f)s_{B}(f)^{*}] (85)
\displaystyle\simeq fBρRe[hA(f)hB(f)+nA(f)nB(f)]\displaystyle\sum_{f\in B_{\rho}}\mathrm{Re}[h_{A}(f)h_{B}(f)^{*}+n_{A}(f)n_{B}(f)^{*}] (86)
\displaystyle\simeq μuρ+fBρRe[nA(f)nB(f)].\displaystyle\braket{\mu_{u}^{\rho}}+\sum_{f\in B_{\rho}}\mathrm{Re}[n_{A}(f)n_{B}(f)^{*}]~{}. (87)

In Eq. (86), the first term comes from the background and can be coherently amplified. On the other hand, the second term is due to the noises and is not amplified because of its incoherence.

Let us calculate the expectation value and the variance of the compressed estimator μuρ\mu_{u}^{\rho}. From Eqs. (32) and (33), the expectation value μuρ\braket{\mu_{u}^{\rho}} is given by

μuρ=\displaystyle\braket{\mu_{u}^{\rho}}= fBρRe[hA(f)hB(f)]\displaystyle\sum_{f\in B_{\rho}}\mathrm{Re}[\braket{h_{A}(f)h_{B}(f)^{*}}] (88)
\displaystyle\sim 8π5Tobsδf(P=T,V,SγuIP(fρ)IP(fρ)\displaystyle\frac{8\pi}{5}T_{obs}\delta f\left(\sum_{P=T,V,S}\gamma^{I_{P}}_{u}(f_{\rho})I_{P}(f_{\rho})\right. (89)
+P=T,VγuWP(fc)WP(fc)).\displaystyle\left.+\sum_{P=T,V}\gamma^{W_{P}}_{u}(f_{c})W_{P}(f_{c})\right)~{}. (90)

The variance is given by the second term in Eq. (86) as

𝒩uρ=\displaystyle\mathcal{N}_{u}^{\rho}= μuρ2μuρ2\displaystyle\braket{\mu_{u}^{\rho 2}}-\braket{\mu_{u}^{\rho}}^{2} (91)
\displaystyle\sim 12ffnA(f)nA(f)nB(f)nB(f)\displaystyle\frac{1}{2}\sum_{f}\sum_{f^{\prime}}\braket{n_{A}(f)n_{A}^{*}(f^{\prime})}\braket{n_{B}(f)n_{B}^{*}(f^{\prime})} (92)
\displaystyle\sim Tobsδf8NA(fρ)NB(fρ)\displaystyle\frac{T_{obs}\delta f}{8}N_{A}(f_{\rho})N_{B}(f_{\rho})~{} (93)

with no noise correlation between different pairs (e.g., between HK and HL). The last line is obtained by substitution of Eq. (75). These expressions show that expectation value μuρ\braket{\mu^{\rho}_{u}} is proportional to the number of the Fourier modes TobsδfT_{obs}\delta f but the variance 𝒩uρ\sqrt{\mathcal{N}^{\rho}_{u}} is proportional to Tobsδf\sqrt{T_{obs}\delta f}. For Tobsδf1T_{obs}\delta f\gg 1, the background signal is relatively amplified to the noise, as expected.

Combining Eqs. (90) and (91), we obtain the SNR of each bin as

SNRuρ2\displaystyle{\rm SNR}_{u}^{\rho 2} =μuρ2𝒩uρ\displaystyle=\frac{\braket{\mu_{u}^{\rho}}^{2}}{\mathcal{N}_{u}^{\rho}} (94)
2(16π5)2TobsδfNA(fρ)NB(fρ)(P=T,V,SγuIP(fρ)IP(fρ)\displaystyle\sim 2\left(\frac{16\pi}{5}\right)^{2}\frac{T_{obs}\delta f}{N_{A}(f_{\rho})N_{B}(f_{\rho})}\left(\sum_{P=T,V,S}\gamma^{I_{P}}_{u}(f_{\rho})I_{P}(f_{\rho})\right. (95)
+P=T,VγuWP(fρ)WP(fρ))2.\displaystyle\left.+\sum_{P=T,V}\gamma^{W_{P}}_{u}(f_{\rho})W_{P}(f_{\rho})\right)^{2}~{}. (96)

Quadratically summing up all the frequency bin, we obtain total SNR for the detector pair uu as

SNRu2\displaystyle{\rm SNR}_{u}^{2} =ρSNRuρ2\displaystyle=\sum_{\rho}{\rm SNR}_{u}^{\rho 2} (97)
=2Tobs(16π5)2\displaystyle=2T_{obs}\left(\frac{16\pi}{5}\right)^{2} (98)
×df(P=T,V,SγuPIP+T,VγuWPWP)2NA(f)NB(f).\displaystyle\times\int df\frac{\left(\sum_{P=T,V,S}\gamma^{P}_{u}I_{P}+\sum_{T,V}\gamma^{W_{P}}_{u}W_{P}\right)^{2}}{N_{A}(f)N_{B}(f)}~{}. (99)

In this paper, we assume that all detectors have the noise spectrum NALN_{\rm AL} identical to the design sensitivity of the advanced LIGO [46] (see Fig. 6 for NAL(f)N_{\rm AL}(f)). Considering the current status of the LVK-network, this assumption looks unrealistic. However, it is virtually difficult for a largely less sensitive detector to make an effective contribution to the network, and we expect that our assumption will eventually become a reasonable approximation. Note that it is, in principle, straightforward to taking into account the difference between detector noise spectra for the rest of this paper. For simplicity, unless otherwise stated, we also assume flat spectra ΩGWQ(f)f0\Omega_{GW}^{Q}(f)\propto f^{0} for the injected backgrounds.

Refer to caption
Figure 6: Noise power spectrum of advanced LIGO, taken from [46]. The spike around 9Hz is due to the resonance of the anti-vibration components.

In the upper right panel of Table 4, we present SNRu for ΩGWIT=108\Omega_{GW}^{I_{T}}=10^{-8}, setting other four spectra at zero. Similarly, in the lower left part, we show SNRu only with the non-vanishing compoent ΩGWWT=108\Omega_{GW}^{W_{T}}=10^{-8} (ignoring the physical requirement |ΩGWWT|ΩGWIT|\Omega_{GW}^{W_{T}}|\leq\Omega_{GW}^{I_{T}}). In the ten detector pairs, the HL pair has the best sensitivity to ITI_{T}, but the worst sensitivity to WTW_{T}. This is due to the small separation angle β=27\beta=27^{\circ} of the HL pair, as pointed out earlier in Sec. II.4. In contrast, the IL pair has the worst sensitivity to ITI_{T} but the best sensitivity to WTW_{T} with the largest separation angle β=128\beta=128^{\circ}.

Table 4: The upper right corresponds to SNRAB{\rm SNR}_{AB} for ΩGWIT=108\Omega^{I_{T}}_{GW}=10^{-8} setting other four spectra at zero. The lower left is only with ΩGWWV=108\Omega^{W_{V}}_{GW}=10^{-8}.
KAGRA LIGO-I LIGO-H LIGO-L Virgo
KAGRA * 2.16 2.42 1.38 4.47
LIGO-I 2.32 * 2.79 0.34 3.27
LIGO-H 3.67 5.07 * 15.4 3.51
LIGO-L 5.09 6.33 1.04 * 3.83
VIRGO 2.28 3.22 2.56 2.07 *

For a background purely made by ITI_{T}, we have the network sensitivity

SNRIT2=2Tobs(16π5)2𝑑fu(γuIT)2IT2NAL2(f).\displaystyle{\rm SNR}_{I_{T}}^{2}=2T_{obs}\left(\frac{16\pi}{5}\right)^{2}\int df\frac{\sum_{u}\left(\gamma^{I_{T}}_{u}\right)^{2}I_{T}^{2}}{N_{\rm AL}^{2}(f)}~{}. (100)

For the HIKLV network and a flat spectrum, we numerically have

SNRIT=19.0(ΩGWIT108)(Tobs3yr)1/2(SNR0),\displaystyle{\rm SNR}_{I_{T}}=19.0\left(\frac{\Omega_{GW}^{I_{T}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2}(\equiv{\rm SNR}_{0})~{}, (101)

which gives the maximum sensitivity to ITI_{T} achieved by the five detectors. In Eq. (101), we introduced the notation SNR0{\rm SNR}_{0} in order to use this result as a reference value in our study below.

V Separation of the five components

As shown in Eq. (90), the expectation value of a single segment μuρ(f)\left\langle\mu_{u}^{\rho}(f)\right\rangle is given as the linear combination of the five spectra Q={IT,IV,IS,WT,WV}Q=\{I_{T},I_{V},I_{S},W_{T},W_{V}\}. For testing alternative gravity theories, we would like to handle them separately. Such a method has been discussed in the literature (ITI_{T} and WTW_{T} in [33, 36], and IT,IV,I_{T},I_{V}, and ISI_{S} in [47]). Its basic strategy is to take the appropriate linear combinations of the cross correlation signals μuρ(f)\mu_{u}^{\rho}(f) and algebraically isolate the background spectra.

To this end, we need at least 5 detector pairs. This can be satisfied by 4 or more detectors, which provide 6 or more pairs (not equal to 5). Thus the spectral decomposition is actually an overdetermined problem.

Our first objective in this section is to present a simple expression for the signal-to noise ratios SNRQρ{\rm SNR}^{\rho}_{Q} after the algebraic spectral decomposition. However, as outlined in Sec. VA, under the orthodox approach, we have a technical difficulty at deriving the simplified expression SNRQρ{\rm SNR}^{\rho}_{Q}. Thus, basically following the arguments in Ref. [36], we provide the desired expression that is not proven in a precise mathematical sense.

V.1 Orthodox Approach

As an example, let us consider the five detector network with 10 data set μuρ\mu_{u}^{\rho} (u=1,,10)(u=1,\cdots,10). Each segment contains the five polarization spectra as in Eq. (90). Using the difference between the ORFs, we can isolate a specific spectrum QQ (e.g., IVI_{V}) by algebraically cancelling other four spectra (e.g., IT,IS,WT,WVI_{T},I_{S},W_{T},W_{V}). Then we obtain the six linear combinations of the original data μuρ\mu_{u}^{\rho}.

In contrast to the original ten data μuρ\mu_{u}^{\rho}, the resultant six combinations have correlated detector noises. We can newly generate six noise orthogonal combinations, as a standard eigenvalue decomposition for the 6×66\times 6 noise matrix. Then, quadratically adding the six orthogonal elements, we obtain the network SNR for the target spectrum QQ. We can formally put

(SNRQρ)2=2Tobsδf(16π5)2Q(f)2XQ(f)NAL2(f).\displaystyle({\rm SNR}_{Q}^{\rho})^{2}=2T_{obs}\delta f\left(\frac{16\pi}{5}\right)^{2}\frac{Q(f)^{2}X_{Q}(f)}{N_{\rm AL}^{2}(f)}. (102)

Here the factor XQ(f)X_{Q}(f) is given by the 50 ORFs, and can be effectively regarded as the square of a compiled ORF.

Unfortunately, following the above line of argument, we could not analytically obtain the simplified symmetrical form for the factor XQ(f)X_{Q}(f) even with Mathematica.

V.2 Alternative Approach

In Ref. [36], a convenient construction scheme was deduced for the factor XQX_{Q}, on the basis of the likelihood study for the multiple spectra (closely related to the Fisher matrix analyses). Here we concisely provide their final expression (see [36] for detail).

We first compose a 5×55\times 5 matrix FF as

FQQu=1npγuQγuQ.\displaystyle F^{QQ^{\prime}}\equiv\sum_{u=1}^{n_{p}}\gamma^{Q}_{u}\gamma^{Q^{\prime}}_{u}~{}. (103)

Next, we take its inverse matrix

ΣF1.\Sigma\equiv F^{-1}~{}. (104)

Then, we presume the following relation for the factor XQX_{Q}

XQ=1ΣQQ(f).X_{Q}=\frac{1}{\Sigma^{QQ}(f)}. (105)

Below, we mention some circumstance evidences for its validity.

For decomposing only two spectra (e.g., ITI_{T} and WTW_{T}), we can analytically confirmed that this relation is actually true for an arbitrary number of detectors. For the five spectral decomposition with ten detector pairs, we numerically generated the 50 ORFs randomly in the range [1,1][-1,1] and evaluated the both sides of Eq. (105) with Mathematica. We repeated this experiments for many times and confirmed equality within numerical accuracy. Note that, with Mathematica, we need much less computational resources at numerical evaluation than at corresponding symbolic processing.

We hereafter use relation (105) and put

(SNRQρ)2=δffZQ(f)(ΩGWQ(f)108)2(Tobs3yr),\displaystyle({\rm SNR}_{Q}^{\rho})^{2}=\frac{\delta f}{f}Z_{Q}(f)\left(\frac{\Omega_{GW}^{Q}(f)}{10^{-8}}\right)^{2}\left(\frac{T_{obs}}{\rm 3yr}\right), (106)

where we defined (e.g., with Eqs. (26) and (102))

ZQ(f)3.7×1082XQ(f)(f1Hz)5(NAL(f)1Hz1)2.Z_{Q}(f)\equiv 3.7\times 10^{-82}{X_{Q}}(f)\left(\frac{f}{\rm 1Hz}\right)^{-5}\left(\frac{N_{AL}(f)}{\rm 1Hz^{-1}}\right)^{-2}~{}. (107)

Here we used H0=70kms1Mpc1H_{0}=70{\rm km~{}s^{-1}~{}Mpc^{-1}}. This function ZQ(f)Z_{Q}(f) shows the contribution of background signals from various frequencies.

After the frequency integral, we obtain

(SNRQ)2=0dffZQ(f)(ΩGWQ(f)108)2(Tobs3yr).\displaystyle({\rm SNR}_{Q})^{2}=\int_{0}^{\infty}\frac{df}{f}Z_{Q}(f)\left(\frac{\Omega_{GW}^{Q}(f)}{10^{-8}}\right)^{2}\left(\frac{T_{obs}}{\rm 3yr}\right). (108)

VI Statistical loss associated with the mode separation

We now examine the matrices FF and Σ\Sigma, in particular the role of their off-diagonal elements.

VI.1 Reduction Factors

For simplicity, we first deal with the two component analysis with the spectra ITI_{T} and QQ^{\prime} (Q=IV,IS(Q^{\prime}=I_{V},I_{S} or WT)W_{T}). The 2×22\times 2 matrix FF is given by

F=\displaystyle F= (u=1npγuITγuITu=1npγuITγuQu=1npγuITγuQu=1npγuQγuQ),\displaystyle\left(\begin{array}[]{cc}\sum_{u=1}^{n_{p}}\gamma_{u}^{I_{T}}\gamma_{u}^{I_{T}}&\sum_{u=1}^{n_{p}}\gamma_{u}^{I_{T}}\gamma_{u}^{Q^{\prime}}\\ \sum_{u=1}^{n_{p}}\gamma_{u}^{I_{T}}\gamma_{u}^{Q^{\prime}}&\sum_{u=1}^{n_{p}}\gamma_{u}^{Q^{\prime}}\gamma_{u}^{Q^{\prime}}\end{array}\right)~{}, (111)

and we have

XIT=1ΣITIT=(1RITQ2)u=1npγuITγuIT,\displaystyle X_{I_{T}}=\frac{1}{\Sigma_{I_{T}I_{T}}}=(1-R_{I_{T}Q^{\prime}}^{2})\sum_{u=1}^{n_{p}}\gamma_{u}^{I_{T}}\gamma_{u}^{I_{T}}~{}, (112)
XQ=1ΣQQ=(1RITQ2)u=1npγuQγuQ.\displaystyle X_{Q^{\prime}}=\frac{1}{\Sigma_{Q^{\prime}Q^{\prime}}}=(1-R_{I_{T}Q^{\prime}}^{2})\sum_{u=1}^{n_{p}}\gamma_{u}^{Q^{\prime}}\gamma_{u}^{Q^{\prime}}~{}. (113)

Here we defined the coefficient RITQR_{I_{T}Q^{\prime}} by

RITQ\displaystyle R_{I_{T}Q^{\prime}}\equiv u=1npγuITγuQu=1nt(γuIT)2u=1np(γuQ)2.\displaystyle\frac{\sum_{u=1}^{n_{p}}\gamma^{I_{T}}_{u}\gamma^{Q^{\prime}}_{u}}{\sqrt{\sum_{u=1}^{n_{t}}\left(\gamma^{I_{T}}_{u}\right)^{2}}\sqrt{\sum_{u=1}^{n_{p}}\left(\gamma^{Q^{\prime}}_{u}\right)^{2}}}~{}. (114)

From Cauchy-Schwartz inequality, we have |RITQ|1|R_{I_{T}Q^{\prime}}|\leq 1 with equality only for two parallel vectors {γuIT}\{\gamma^{I_{T}}_{u}\} and {γuQ}\{\gamma^{Q^{\prime}}_{u}\}. The coefficient RITQR_{I_{T}Q^{\prime}} represents the correlation between the two spectra and reduces SNRs after the spectral decomposition through the factor (1RITQ2)(1-R_{I_{T}Q^{\prime}}^{2}) (see Eqs. (102) and (112)). This factor shows the statistical loss associated with the decomposition.

So far, we discussed two component decomposition. When the number nQn_{Q} of the target spectral components is larger than two (nQ>2n_{Q}>2), we can similarly define the reduction factor 1RQi21-R_{Q_{i}}^{2} (i=1,,nQi=1,\cdots,n_{Q}) by

1RQi2=XQi(f)uγuQiγuQi=1(F1)QiQiFQiQi.\displaystyle 1-R_{Q_{i}}^{2}=\frac{X_{Q_{i}}(f)}{\sum_{u}\gamma_{u}^{Q_{i}}\gamma_{u}^{Q_{i}}}=\frac{1}{(F^{-1})^{Q_{i}Q_{i}}F^{Q_{i}Q_{i}}}~{}. (115)

Note that we omitted the subscripts other than the component of interest for the notational simplicity. If the vectors {γuQi}\{\gamma_{u}^{Q_{i}}\} (i=1,,nQi=1,\cdots,n_{Q}) are close to linearly dependent, the matrix FF becomes nearly singular, and we could have |(F1)QiQiFQiQi|1|(F^{-1})^{Q_{i}Q_{i}}F^{Q_{i}Q_{i}}|\gg 1, resulting in a large signal loss 1RQi211-R_{Q_{i}}^{2}\ll 1. In this relation, our two new findings in Sec. II could play interesting roles, as explained in the next subsection.

VI.2 Numerical Results

In Fig. 7, we show the reduction factor (1RIT2)(1-R_{I_{T}}^{2}) for the two component models {IT,IV}\{I_{T},I_{V}\} (upper) and {IT,IS}\{I_{T},I_{S}\} (lower). As shown in Eq. (68), we have the degeneracy limf0{γuIT}={γuIV}={γuIS}\lim_{f\to 0}\{\gamma_{u}^{I_{T}}\}=\{\gamma_{u}^{I_{V}}\}=\{\gamma_{u}^{I_{S}}\} and need the sub-leading correction O(f2)O(f^{2}) to decompose the two spectra. We thus have a significant suppression (1RIT2)0.1(1-R_{I_{T}}^{2})\lesssim 0.1 at f10f\lesssim 10Hz.

Refer to caption
Refer to caption
Figure 7: The reduction factors 1RITIV21-R_{I_{T}I_{V}}^{2} (upper) and 1RITIS21-R_{I_{T}I_{S}}^{2} (lower) respectively for the two component analyses {IT,IV}\{I_{T},I_{V}\} and {IT,IS}\{I_{T},I_{S}\}.
Refer to caption
Figure 8: The reduction factors 1RWTWV21-R_{W_{T}W_{V}}^{2} for the hypothetical two component analysis {WT,WV}\{W_{T},W_{V}\}.
Refer to caption
Figure 9: The reduction factors 1RTT21-R_{T_{T}}^{2} for the three and five component analyses.
Refer to caption
Figure 10: The reduction factors 1RITWT21-R_{I_{T}W_{T}}^{2} for the two component analysis {IT,WT}\{I_{T},W_{T}\}.

In Fig. 8, we examined the hypothetical case for decomposing the two odd spectra {WT,WV}\{W_{T},W_{V}\}. Their ORFs are parallel at the low frequency limit and nearly parallel around the anathematic frequency 13Hz. We thus have the siginificant signal reduction below 13Hz, as in Fig. 8.

Next we move to examine the decomposition of more than two spectra nQ>2n_{Q}>2. In Fig. 9, we show the reduction factor (1RIT2)(1-R_{I_{T}}^{2}) at separating the three even spectra {IT,IV,IS}\{I_{T},I_{V},I_{S}\}. In contrast to the two component cases in Fig. 7, the strong suppression 1RIT20.11-R_{I_{T}}^{2}\lesssim 0.1 continues up to 2020Hz. This is because the trinity degeneracy (71) works still at the sub-leading order O(f2)O(f^{2}). We thus need the higher corrections O(f4)O(f^{4}) to isolate the three spectra. In fact, even for a detector network not tangential to a sphere, we still have detF=O(f4)\det F=O(f^{4}) for the 3×33\times 3 matrix FF of the three even spectra {IT,IV,IS}\{I_{T},I_{V},I_{S}\}.

Note that the LIGO-India plays a key role for the usage of the higher order terms O(f4)O(f^{4}) (or more appropriately O(y4)O(y^{4}) for the perturbative expansion). In Fig. 9, we can clearly see the resulting improvement around 20-40Hz. Here the mechanism around Eq. (73) works efficiently, in particular, with the HI and IL pairs.

In Fig. 10, we present the result for the two tensorial spectra {IT,WT}\{I_{T},W_{T}\}. Since their ORFs {γuIT}\{\gamma_{u}^{I_{T}}\} and {γuWT}\{\gamma_{u}^{W_{T}}\} are generally not parallel, the reduction is not significant. If we use the HIKLV pair, the reduction factor is no less than 0.8.

VII Signal to noise ratio

VII.1 Results for the HIKLV Network

Now we discuss the signal-to-noise ratios SNRQ{\rm SNR}_{Q} after the spectral decomposition and the associated frequency profiles ZQ(f)Z_{Q}(f) defined in Eq. (107). We start with the results for the HIKLV network and flat spectra ΩGWQ=const\Omega_{GW}^{Q}={\rm const}.

In Fig. 11, we show the profile ZIT(f)Z_{I_{T}}(f) for ITI_{T}. The sharp dip around 10Hz is caused by the noise spike in Fig. 5. The uppermost blue line shows the result for the simplest case only with ITI_{T} (no reduction factor). Its peak is around 25Hz with the integrated value SNRIT{\rm SNR}_{I_{T}} (see Eq. (101))

SNR0=19.0(ΩGWIT108)(Tobs3yr)1/2.{\rm SNR}_{0}=19.0\left(\frac{\Omega_{GW}^{I_{T}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2}. (116)

We use this expression to normalize the signals SNRQ{\rm SNR}_{Q} in different settings, as presented in Tables V and VI. In Fig. 11, the four lines other than the blue one show the profiles ZIT(f)Z_{I_{T}}(f) after decomposing multiple spectra. Their fractional differences from the blue lines represent the corresponding reduction factor (1RIT2)(1-R_{I_{T}}^{2}).

At the decomposition of ITI_{T} and WTW_{T} (dashed orange line in Fig. 12), the statistical loss is inconspicuous with the total value SNRIT/SNR0=0.99{\rm SNR}_{I_{T}}/{\rm SNR}_{0}=0.99 (see Table V). However, we need to pay a significant cost to isolate the three even spectra ITI_{T}, IVI_{V} and ISI_{S}. The total signal decreases down to SNRIT/SNR0=0.47{\rm SNR}_{I_{T}}/{\rm SNR}_{0}=0.47 and the peak of the profile ZIT(f)Z_{I_{T}}(f) moves up to 40\sim 40Hz.

In Figure 12 and Table VI, we show the results for the odd tensor spectrum WTW_{T}. Similarly to Fig. 11, we can isolate it from ITI_{T} with almost no loss of the integrated signal SNRWT{\rm SNR}_{W_{T}} (see Table VI). When we separate WTW_{T} and WVW_{V}, the anathematic frequency 13Hz clearly appears, as shown by the green and red lines, and the function ZWT(f)Z_{W_{T}}(f) is significantly suppressed below 20\sim 20Hz. In contrast to ZIT(f)Z_{I_{T}}(f), the peak of the profile ZWT(f)Z_{W_{T}}(f) stays around 25Hz.

Refer to caption
Figure 11: The factor ZIT(f)Z_{I_{T}}(f) showing the signal strength defined in Eq. (107) for the HIKLV network. The blue and orange curves are nearly overlapped. The ratio between the blue and other curves corresponds to the reduction factor 1RIT21-R_{I_{T}}^{2} due to the signal correlation. The sharp dip around 9Hz is due to the noise spectrum NAL(f)N_{AL}(f).
Refer to caption
Figure 12: The factor ZWT(f)Z_{W_{T}}(f) showing the signal strength defined in Eq. (107) for the HIKLV network. The green and the red curves have the sharp dips around 13Hz.
Refer to caption
Refer to caption
Figure 13: The top and bottom panels respectively show ZIT(f)Z_{I_{T}}(f) and ZIT(f)Z_{I_{T}}(f) for three networks. The dotted lines in the upper panel are proportional to f2f^{2} and f2f^{-2}.

VII.2 LIGO-India

Next we discuss the impacts of adding LIGO-India to the detector network. As shown in Table V, for the single spectral search ITI_{T}, LIGO-India increases the total signal SNRITSNR_{I_{T}} by only 3%. However, together with KAGRA, it makes a notable contribution to improve the sensitivity to the odd spectra WTW_{T} (see Table VI). In addition, as explained earlier, LIGO-India also helps us to use the higher order terms O(f4)O(f^{4}) for decomposing the three even spectra. For the five spectral search, we can double both SNRIT{\rm SNR}_{I_{T}} and SNRWT{\rm SNR}_{W_{T}} by adding the LIGO-India detector.

VII.3 Power-law Models

So far, we have assumed that the background has flat spectra ΩGWQ=const\Omega_{GW}^{Q}={\rm const}. Now, we briefly discuss a power-law form ΩGWQfα\Omega_{GW}^{Q}\propto f^{\alpha} in the frequency regime in interest.

The integrated signal SNRQ{\rm SNR}_{Q} in Eq. (108) has the dominant contribution around the frequency where the function ZQ(f)Z_{Q}(f) is tangential to a curve f2αf^{-2\alpha}. As deduced from Fig. 13, for the five spectral decompositions with the HIKLV network, the tangential frequencies are 40Hz for ITI_{T} and 25Hz for WTW_{T}, as long as the index α\alpha is in the range [1,1][-1,1]. Therefore, the total signals are very roughly given as

SNRIT19×0.4(ΩGWIT(40Hz)108)(Tobs3yr)1/2\displaystyle{\rm SNR}_{I_{T}}\sim 19\times 0.4\left(\frac{\Omega_{GW}^{I_{T}}({\rm 40Hz})}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (117)
SNRWT19×0.39(ΩGWWT(25Hz)108)(Tobs3yr)1/2.\displaystyle{\rm SNR}_{W_{T}}\sim 19\times 0.39\left(\frac{\Omega_{GW}^{W_{T}}({\rm 25Hz})}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2}. (118)
Table 5: Ratio SNRIT/SNR0{\rm SNR}_{I_{T}}/{\rm SNR}_{0} after the spectral isolation. All five components of LHV is missing since LHV has only three independent detector pairs. We assumed a flat spectrum ΩGWIT=const\Omega_{GW}^{I_{T}}={\rm const}.
background components KILHV KLHV LHV
ITI_{T} only 1 0.97 0.91
IT,WTI_{T},W_{T} 0.99 0.96 0.91
IT,IVI_{T},I_{V} 0.63 0.57 0.47
IT,ISI_{T},I_{S} 0.89 0.82 0.74
IT,IV,ISI_{T},I_{V},I_{S} 0.47 0.33 0.22
All five 0.40 0.20 *
Table 6: Ratio SNRWT/SNR0{\rm SNR}_{W_{T}}/{\rm SNR}_{0} after the spectral isolation. We assume flat spectra and omit the factor ΩGWWT/ΩGWIT\Omega_{GW}^{W_{T}}/\Omega_{GW}^{I_{T}} for simplicity.
background components KILHV KLHV LHV
WTW_{T} only 0.62 0.40 0.18
WT,ITW_{T},I_{T} 0.62 0.39 0.18
WT,WVW_{T},W_{V} 0.43 0.20 0.06
All five 0.39 0.17 *

VIII Summary

In this paper, we studied the prospects for the polarizational study of isotropic stochastic gravitational wave backgrounds by correlating second generation detectors. In the long-wave approximation, the backgrounds are generally characterized by the five spectra IT,V,SI_{T,V,S} and WT,VW_{T,V}. The modes other than ITI_{T} can appear in modified theories of gravity.

For correlation analysis, the ORFs play key roles. In this paper, we newly identified two simple relations behind them. The first one is the trinity degeneracy (71) between the three even ORFs at the sub-leading order O(f2)O(f^{2}). The second one is the degeneracy between the two odd ORFs around the specific frequency 13Hz.

For each detector pair, the correlation product is given as a linear combination of the five spectra. To closely examine theories of gravitation, we desire to separate the five spectra clearly. We thus examined their algebraic decomposition using the difference between the involved ORFs. Here we generally need to handle an over-determined problem. By extending an analytic framework in the literature, we derived the formal expression (108) for the optimal SNRs after the spectral decomposition.

Then, assuming an identical noise curve for the five detectors and flat background spectra, we discussed the statistical loss of sensitivities accompanied by the decomposition. This loss is closely related to the off-diagonal elements of the matrix FQQi=1npγuQγuQF^{QQ^{\prime}}\propto\sum_{i=1}^{n_{p}}\gamma_{u}^{Q}\gamma_{u}^{Q^{\prime}}.

In this context, our two findings are quite useful for following the singular behaviors at the decomposition. On the one hand, when simultaneously dealing with the three even spectra, due to the higher order degeneracy of their ORFs, we have a large signal reduction below 20Hz, unlike the two spectral decomposition (such as ITIVI_{T}-I_{V} and ITISI_{T}-I_{S}). On the other hand, it is very difficult to separate the two odd spectra below 20\sim 20Hz, including the anathematic frequency 13Hz. Given the structure of the covariance matrix FF, these limitations will also appear in the likelihood or Fisher matrix analyses.

We also discussed the advantage of adding the LIGO-India detector to the ground-based detector network. As shown in Tables V and VI, it can largely increase the sensitivities to the odd spectra and will also help us to decompose multiple spectra. Here the HI and LI pairs are particularly useful with the large separation angles β\beta.

In this paper, we have mainly considered the second generation ground-based detectors. However, our method is general enough to be straightforwardly applied to the third generation ground-based detectors (such as ET [48] and CE [49], see also  [50]) and partially to space borne detectors (LISA [39], TAIJI [40], and TianQin [41]). The former will cover a lower frequency regime than that of the second generation ones and will be more severely affected by the limitations associated with our two findings.

Acknowledgements.
We would like to thank M. Ando and S. Bose for valuable comments. This work is supported by JSPS Kakenhi Grant-in-Aid for Scientific Research (Nos. 17H06358 and 19K03870). HO is supported by Grant-in-Aid for JSPS Fellows JP22J14159.

Appendix A optimal SNR for the ground-based detectors

Assuming the flat spectrum of the background and using Eq. (108), we can evaluate the SNR for each spectra after the decomposition. As a reference, we provide numerical results for the five spectral components. For the HKLV-network, we obtain

SNRIT\displaystyle{\rm SNR}_{I_{T}} =3.94(ΩGWIT108)(Tobs3yr)1/2\displaystyle=3.94\left(\frac{\Omega_{GW}^{I_{T}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (119)
SNRIV\displaystyle{\rm SNR}_{I_{V}} =2.75(ΩGWIV108)(Tobs3yr)1/2\displaystyle=2.75\left(\frac{\Omega_{GW}^{I_{V}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (120)
SNRIS\displaystyle{\rm SNR}_{I_{S}} =6.81(ΩGWIS108)(Tobs3yr)1/2\displaystyle=6.81\left(\frac{\Omega_{GW}^{I_{S}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (121)
SNRWT\displaystyle{\rm SNR}_{W_{T}} =3.14(ΩGWWT108)(Tobs3yr)1/2\displaystyle=3.14\left(\frac{\Omega_{GW}^{W_{T}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (122)
SNRWV\displaystyle{\rm SNR}_{W_{V}} =4.07(ΩGWWV108)(Tobs3yr)1/2.\displaystyle=4.07\left(\frac{\Omega_{GW}^{W_{V}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2}. (123)

For the HIKLV-network, we have

SNRIT\displaystyle{\rm SNR}_{I_{T}} =7.53(ΩGWIT108)(Tobs3yr)1/2\displaystyle=7.53\left(\frac{\Omega_{GW}^{I_{T}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (124)
SNRIV\displaystyle{\rm SNR}_{I_{V}} =6.14(ΩGWIV108)(Tobs3yr)1/2\displaystyle=6.14\left(\frac{\Omega_{GW}^{I_{V}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (125)
SNRIS\displaystyle{\rm SNR}_{I_{S}} =9.74(ΩGWIS108)(Tobs3yr)1/2\displaystyle=9.74\left(\frac{\Omega_{GW}^{I_{S}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (126)
SNRWT\displaystyle{\rm SNR}_{W_{T}} =7.50(ΩGWWT108)(Tobs3yr)1/2\displaystyle=7.50\left(\frac{\Omega_{GW}^{W_{T}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2} (127)
SNRWV\displaystyle{\rm SNR}_{W_{V}} =8.27(ΩGWWV108)(Tobs3yr)1/2.\displaystyle=8.27\left(\frac{\Omega_{GW}^{W_{V}}}{10^{-8}}\right)\left(\frac{T_{obs}}{3{\rm yr}}\right)^{1/2}. (128)

References