ebook img

Lensing Effects on Gravitational Waves in a Clumpy Universe -Effects of Inhomogeneity on the Distance-Redshift Relation- PDF

1.6 MB·
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Lensing Effects on Gravitational Waves in a Clumpy Universe -Effects of Inhomogeneity on the Distance-Redshift Relation-

Department of Mathematics and Physics OCU-PHYS-244 Osaka City University AP-GR-32 Lensing Effects on Gravitational Waves in a Clumpy Universe —Effects of Inhomogeneity on the Distance-Redshift Relation— 7 0 Chul-Moon Yoo1, Ken-ichi Nakao2and Hiroshi Kozaki3 0 2 Department of Mathematics and Physics, Graduate School of Science, Osaka City n a University, Osaka 558-8585, Japan J 5 Ryuichi Takahashi4 3 v Division of Theoretical Astronomy, National Astronomical Observatory of Japan, Mitaka, 3 2 Tokyo 181-8588, Japan 1 4 0 ABSTRACT 6 0 / h The distance-redshift relation determined by means of gravitational waves in p - the clumpy universe is simulated numerically by taking into account the effects o r of gravitational lensing. It is assumed that all of the matter in the universe t s takes the form of randomly distributed point masses, each of which has the a : v identical mass ML. Calculations are carried out in two extreme cases: λ ≫ Xi GML/c2 and λ GML/c2, where λ denotes the wavelength of gravitational ≪ r waves. In the first case, the distance-redshift relation for the fully homogeneous a and isotropic universe is reproduced with a small distance dispersion, whereas in the second case, the distance dispersion is larger. This result suggests that we might obtain information about the typical mass of lens objects through the distance-redshift relation gleaned through observation of gravitational waves of various wavelengths. In this paper, we show how to set limitations on the mass M through the observation of gravitational waves in the clumpy universe model L described above. Subject headings: gravitational lensing – gravitational waves – dark matter – distance scale – 2 – 1. Introduction Owing to recent developments of observational techniques, it is expected that detection of gravitational waves will be achieved in the near future. Laser interferometer gravitational wave detectors, such as the Tokyo Advanced Medium-Scale Antenna (TAMA-300), the Laser Interferometer Gravitational-Wave Observatory (LIGO), Variability of Irradiance and Grav- ity Oscillations (VIRGO) and GEO-600, are in operation, and other projects, such as the Large-scale Cryogenic Gravitational wave Telescope (LCGT), LIGO2, the Laser Interfer- ometer Space Antenna (LISA), the Decihertz Interferometer Gravitational Wave Observa- tory (DECIGO; Seto et al. (2001)), and the Big Bang Observer (BBO), are in the planning stage. In preparation for future breakthroughs associated with these projects, it is useful to discuss the information provided by gravitational wave data. Here we focus on inhomo- geneities of our universe that can be investigated with gravitational wave observations. Many cosmological observations suggest that our universe is globally homogeneous and isotropic, in other words, that the time evolution of the global aspect is well approximated by the Friedmann-Lemaˆıtre(FL) cosmological model. However, our universe is locally in- homogeneous. The effects of these inhomogeneities on the evolution of our universe or on cosmological observations draw our attention as one of the most important issues in cosmol- ogy. However, it is difficult to directly observe the inhomogeneities with optical observations because most of the matter in our universe is not luminous. One useful means of obtaining information about aspects of inhomogeneities is to examine gravitational lensing. Gravita- tional and electromagnetic waves are subject to gravitational lensing effects caused by inho- mogeneously distributed matter around their trajectories. Therefore, we may find features of inhomogeneities through both observational and theoretical studies of the gravitational lensing effects in astronomically significant situations. There are various candidates for the dark matter such as weakly interacting massive particles (WIMPs) and massive compact halo objects(MACHOs). Several observations have been undertaken to constrain the mass density of compact objects, Ω ; direct searches for CO MACHOs in the Milky Way have been performed by the MACHO and EROS collaborations through microlensing surveys. The MACHO group (Alcock et al. 2000) group concluded that the most likely halo fraction in form of compact objects with masses in the range 0.1- 1E-mail:c m [email protected] 2E-mail:[email protected] 3E-mail:[email protected] 4E-mail:[email protected] – 3 – 1M is of about 20%. The EROS team concluded that the objects in the mass range from ⊙ 2 10−7M to 1M cannot contribute more than 25% of the total halo (Afonso et al. 2003). ⊙ ⊙ × However, the universal fraction of macroscopic dark matter could be significantly different from these local estimates. Millilensing tests of gamma-ray bursts derive a limit on Ω of CO Ω < 0.1 in the mass range from 105 to 109M (Nemiroff et al. 2001). Multiple imaging CO ⊙ searches in compact radio sources derive the limit on Ω as Ω . 0.013 in the mass range CO CO from 106 to 108M (Wilkinson et al. 2001). The mass range of these observational tests ⊙ are limited, and thus other methods to investigate the mass fraction of macroscopic compact objects in wider mass range are needed. In this paper, we show that it is possible to extract information about the properties of macroscopic compact objects fromtheobservationaldataofgravitationalwaves byanalyzing the gravitational lensing effects due to these compact objects. For this purpose, we consider an idealized model of the inhomogeneous universe which has the following properties. First, this model is a globally FL universe. Second, all of the matter takes the form of point masses, each of which has the identical mass M . Finally, the point masses are uniformly L distributed. One of the main targets of gravitational-wave astronomy is binary compact objects. Their gravitational waves have much longer wavelengths (λ) than the optical, and further- more, these are coherent. In typical optical observations, the wavelength (λ 1µm) might ∼ be much shorter than the Schwarzschild radii of lens objects.5 Thus, we usually analyze the gravitational lensing effects on electromagnetic waves by using geometrical optics. In contrast, we need wave optics for the gravitational lensing of gravitational waves, since the wavelength of gravitational waves may be comparable to or longer than the Schwarzschild radii of lens objects. For example, the wave-band of LISA is typically 1011 1014 cm; a point − mass of 105 108M has a Schwarzschild radius almost equal to the wavelengths within ⊙ − that wave-band. When we consider gravitational lensing effects on the gravitational waves with GM /c2 . λ, we have to take the wave effect (Nakamura 1998) into account. In L fact, remarkable differences between the extreme cases λ GM /c2 and λ GM /c2 have L L ≫ ≪ already been reported (Takahashi & Nakamura 2003). In order to obtain information about uniformly distributed compact objects by investi- gatingthegravitationallensingeffectsinobservationaldata,wefocusontherelationbetween the distance from the observer to the source of the gravitational waves and its redshift. The 5It does not conflict with any observationalresult that theremight be lens objects of Schwarzschildradii muchsmallerthan 1µm. However,forexample, supernovaewhichmightbe the smallestand yetverybright opticalsources,arelargerthantheEinsteinradiiofsuchsmall-masslensobjects,andthusthelensingeffects due to those will be negligible. – 4 – distance is determined using the information contained in the amplitudes of the gravita- tional waves from the so-called standard sirens (Holz & Hughes 2005). Although there are several similar analyses using the optical observation of Type Ia supernovae (Holz & Linder 2005), gravitational lensing effects on gravitational waves can add new information about the properties of the lensing compact objects to that obtained by optical observations, by virtue of the wave effect, through which we can gain an understanding of the typical mass and number density. Hereweassumethattheredshiftofeachsourceisindependentlygivenbytheobservation of the electromagnetic counterpart (Kocsis et al. 2006) or the waveform of the gravitational waves (Markovic1993). Theluminositydistanced fromtheobserver tothesourceisgivenby l the waveform of the gravitational waves, as follows (Schutz 1986). Since the orbit of binary will be quickly circularized by a gravitational radiation reaction (Peters & Mathews 1963; Peters 1964), the effect of ellipticity is negligible when the emitted gravitational radiation becomes so strong that we can observe it. Then, the two wave modes emitted from a binary of two compact objects with masses m and m are given by 1 2 2 5/3 (πf)2/3cos(2πft) h = Mchirp (cos2θ +1), (1) + d l 2 5/3 (πf)2/3sin(2πft) h = Mchirp 2cosθ, (2) × d l where f and θ are the frequency and the angle between the angular momentum vector of the binary and the line-of-sight, respectively, and is the redshifted chirp mass, defined by chirp M (m m )3/5 1 2 = (1+z)M = (1+z) . (3) Mchirp chirp (m +m )1/5 1 2 Assuming a circular orbit, we have df 96π8/3 = f11/3 5/3 . (4) dt 5 Mchirp Eliminating and θ from equations (1), (2), and (4), we can obtain d if there is no chirp l M gravitational lensing effect on the gravitational waves. However, if the gravitational waves are gravitationally lensed, their wave forms are changed. The gravitational lensing effects may lead to an incorrect estimation of d , and thus we have to know how the gravitational l lensing changes the waveforms of the gravitational waves. This paper is organized as follows. In 2, the basic theory of gravitational lensing is § introduced. We discuss the lensing probability in 3. Then we show the calculation method § and results for the distance-redshift relation in 4 and 5. Finally, 6 is devoted to the §§ § conclusions and summary. Throughout the paper, we adopt the unit of G = c = 1. – 5 – 2. Review of gravitational lensing In this section, we introduce basic equations for gravitational lensing. 2.1. Geometrical optics In the case of a point-mass lens with mass M , the bending angle vector αˆ(ξ) is given L by (Schneider et al. 1992) ξ αˆ(ξ) = 4M , (5) L ξ 2 | | where ξ is the impact vector (see Fig. 1). For convenience, we consider the “straight” line A from a source to the observer and define the intersection point of A with the lens plane as the origin of the lens position ζ and the ray position γ. We note that, in general, the vector ξ is not orthogonal to the line A. However, if the geometrically thin lens approximation is valid, we can regard the vectors ξ, ζ, and γ as parallel to each other (Schneider et al. 1992). We define D , D , and D as the angular diameter distances from the observer to the S L LS source, from the observer to the lens, and from the lens to the source, respectively. Using simple trigonometry (see Fig. 1), we find the relation between the source position η and ξ, D η = Sξ D αˆ(ξ). (6) LS D − L Equation (6) becomes D 0 = Sγ D αˆ(γ ζ) (7) LS D − − L in terms of ζ and γ. Here we introduce the dimensionless vectors defined by ξ η x = and l = , (8) ξ η 0 0 where ξ is the Einstein radius given by 0 D D L LS ξ = 4M (9) 0 L D r S and D S η = ξ . (10) 0 0 D L – 6 – Fig. 1.— Geometry of gravitational lensing by a point-mass lens. The impact vector ξ represents the relative position of a light ray on the lens plane to the lens position, and αˆ is a vector whose norm is equal to the bending angle. From Figure 1, we find the relation ζ l = . (11) −ξ 0 The source will have two images at position (x ,x ). Substituting equation (5) into equa- + − tion (6) we find 1 x = √l2 +4 l , (12) ± 2 ± (cid:16) (cid:17) where x = x and l = l . The total magnification factor µ is given by ± ± | | | | ∂l ∂l l2 +2 µ = µ + µ = det + det = . (13) | +| | −| ∂x ∂x l√l2 +4 (cid:12) (cid:18) +(cid:19)(cid:12) (cid:12) (cid:18) −(cid:19)(cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) 2.2. Wave optics in gravitational lensing due to point mass lens For simplicity, let us consider gravitational waves that propagate in asymptotically flat spacetime with the metric ds2 = (1+2U)dt2 +(1 2U)dr2, (14) − − where U is the gravitational potential and we assume U 1. If the wavelength of the | | ≪ gravitational waves is much smaller than the typical curvature radius of the background – 7 – spacetime, the equation for the gravitational waves is equivalent to that of scalar waves. Assuming monochromatic waves of an angular frequency ω, we have (Peters 1974) 2 +ω2 φ = 4ω2Uφ. (15) ∇ (cid:0) (cid:1) For convenience, we introduce the amplification factor F defined by φL(ω) F = , (16) φ(ω) where φ is the plane wave with no lensing effects, and φL is the wave that is undergoing lensing effects. Here note that φL is no longer a plane wave, although it was initially. In the caseofthepoint-masslenswithmassM ,thegravitationalpotentialisgivenbyU = M /r. L L − Then we have the solution of equation (15) in the form πw w w i i i F(ω;y) = exp +i ln Γ 1 w F w,1; wy2 , (17) 1 1 4 2 2 − 2 2 2 (cid:18) (cid:19) (cid:18) (cid:19) h (cid:16) (cid:17)i where Γ and F are the gamma function andthe confluent hypergeometric function, respec- 1 1 tively, and w := 4ωM (Peters 1974). Since U is much larger than the Hubble parameter L |∇ | of the universe, by replacing ω with ω(1 + z ), we can use equation (17) in cosmological L situations. 2.3. The Dyer-Roeder distance Intheclumpyuniversemodel,thelocalgeometryisinhomogeneous, butitsglobalaspect might be well described by the FL universe whose metric is given by dR2 ds2 = dT2 +a2(T) +R2dΩ2 , (18) − 1+KR2 (cid:18) (cid:19) where K = 1, 0, and 1, and dΩ2 is the round metric. However, the global aspect of the − clumpy universe is a vague notion since we do not have a definite mathematical prescription to identify the global aspect of the locally inhomogeneous universe with the FL universe model. Therefore, the above FL universe is a fictitious background universe; hereafter, we will refer to this fictitious geometry (universe) as the “global geometry” (or universe). We assume that the global geometry is well approximated by that of the FL universe filled with dust and with a cosmological constant. In the clumpy universe model, the distances D , D , and D in equation (6) or equa- L S LS tion (7) are replaced by the Dyer-Roeder (DR) distances (Kantowski 1969; Dyer & Roeder – 8 – 1973), which are the observed angular diameter distances if the gravitational waves prop- agate far from the point masses. The DR distance with smoothness parameter α˜ = 0 (Schneider et al. 1992) in the global universe defined above is given by 1+z z2 1 dz 1 D (z ,z ) = , (19) DR 1 2 H (1+z)2 Ω (1+z)3 Ω (1+z)2 +Ω 0 Zz1 m0 − K0 Λ0 where z and z are the redshift of the obserpver and source, and Ω , H , and Ω are the 1 2 m0 0 Λ0 present valuesofthetotaldensity parameter, Hubbleparameterandnormalizedcosmological constant in the global universe, respectively, and Ω = Ω + Ω 1. Hereafter, the K0 m0 Λ0 − subscript 0 means the present value. Some further discussions of the properties of the DR distance are presented in Linder (1988), Seitz & Schneider (1994) , Kantowski (1998), and Sereno et al. (2001). 3. The distribution of point masses In order for the clumpy universe to be the same as a homogeneous andisotropic universe from a global perspective, the point masses must be distributed uniformly. However, since there are no isometries of homogeneity and isotropy in the clumpy universe, the notion of “uniform distribution” cannot be introduced in a strict sense. Thus, we assume a uniform distribution of the point masses with respect to the global geometry represented by equa- tion (18) in a manner consistent with the mass density of this global universe. This universe modeliscalled“theon-averageFriedmannuniverse.” Theassumptionsinthisuniverse model are discussed in Seitz et al. (1994). The comoving number density ρ of the point masses M is given by n L a3ρ 3Ω H2 ρ = = m0 0, (20) n M 8πM L L where ρ is the average mass density and we have set a = 1. We consider a past light cone in 0 the global universe. The vertex of this light cone is at the observer, at R = 0 and z = 0, and is parametrized by the redshift z of its null geodesic generator. Then the comoving volume of a spherical shell bounded by R(z) and R = R(z +∆z) on this light cone is given by dR ∆V = 4πR2 ∆z. (21) dz Therefore, the number of point masses in this shell is given by 3Ω H2 dR ∆N = ρ ∆V = m0 0R2 ∆z, (22) n 2M dz L – 9 – where dR 1 1+H2Ω (1+z)2D2(z) = 0 K0 F , (23) dz H0sΩm0(1+z)3 ΩK0(1+z)2 +ΩΛ0 − whereD (z)istheangulardiameter distancefromthesourceoftheredshift z totheobserver F in the global universe, which is expressed as 1 sinX(z) for Ω > 0 K0 H (1+z) √Ω 0 K0  1 D (z) =  Y(z) for Ω = 0 , (24) F  H (1+z) K0  0  1 sinhX(z) for Ω < 0 K0 H (1+z) √ Ω  0 − K0  where  z dz′ X(z) = Ω , (25) K0 | | Ω (1+z′)3 Ω (1+z′)2 +Ω Z0 m0 − K0 Λ0 p z dz′ p Y(z) = . (26) Ω (1+z′)3 Ω (1+z′)2 +Ω Z0 m0 − K0 Λ0 The average number of point masses in the region y < ζ/ξ < y +∆y of this shell is given p 0 by 2πξ2y∆y p(y,z)∆y = 0 ∆N. (27) 4πa2R2 Substituting equations (22) and (23) into equation (27) and using equation (9), we have 1+H2Ω (1+z)2D2(z) p(y,z)∆y = 3H Ω (1+z)2 0 K0 F 0 m0 sΩm0(1+z)3 ΩK0(1+z)2 +ΩΛ0 − D (0,z)D (z,z ) DR DR S y∆y∆z, (28) × D (0,z ) DR S where z is the redshift of the source and we have replaced each of distances in equation (9) S with the DR distance. We find that p(y,z) does not depend on M , and therefore the lensing L probability for a given source does not depend on M . L We note that the assumption of the distribution of point masses given in equation (28) does not necessarily guarantee equality between the magnification in the global universe defined by 2 D (0,z) DR µ := (29) F D (z) (cid:18) F (cid:19) and the average magnification in the clumpy universe. Sevral previous works have discussed this issue (Weinberg 1976; Ellis et al. 1998; Claudel 2000; Rose 2001; Kibble & Lieu 2005). – 10 – 4. The long-wavelength case, λ M L ≫ In the case of λ M , the amplification factor F(ω(1 + z );y) is almost equal to L L ≫ unity (see Fig. 2) because of diffraction effects (Nakamura 1998), and we have φ φ L | − | = F(ω(1+z );y) 1 1. (30) L φ | − | ≪ | | Thus we assume that a wave φ can be treated as a spherical wave whose center is at the source during all stages of the propagation. Let us consider the situation in which N point masses are located at (z ,y ), i = 1,...,N,, where i is the number assigned to each point i i mass in order of distance fromthe observer. If our assumption is valid, the wave φ is changed whenever it propagates near each point mass as N φ F φ F F φ F φ, (31) N N N−1 i → → → ··· → i=1 Y where F = F (ω(1+z );y ). (32) i i i Thus the observed amplitude φ is given by obs | | N φ = F φ . (33) obs i | | | || | i=1 Y Then, the total amplification F is given by total | | N F = F . (34) total i | | | | i=1 Y We randomly distribute point masses so that the distribution of those is consistent with equation (28). First, we divide the spherical region z < z into N concentric spherical S shells, each of which is bounded by two spheres, z = z ∆z/2 and z = z + ∆z/2. We i i − take into account only the nearest lens in each shell. As described below, this procedure is valid for large numbers N, i.e., small ∆z. We find from equation (28) that in the ith shell z ∆z/2 < z < z + ∆z/2, there is one point mass within the region y Y on average, i i i − ≤ where Y is defined by i Yi p(y,z )dy = 1. (35) i Z0 Then we randomly put a point mass within the region y Y in the ith shell. Here it is i ≤ worth noting that Y is an increasing function of N, so by setting N large, we can take into i

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.