<<

Sensitivity curves for searches for gravitational-wave backgrounds

Eric Thrane1, a and Joseph D. Romano2, b 1LIGO Laboratory, California Institute of Technology, MS 100-36, Pasadena, California 91125, USA 2Department of Physics and Astronomy and Center for Gravitational-Wave Astronomy, University of Texas at Brownsville, Texas 78520, USA (Dated: October 19, 2013) We propose a graphical representation of detector sensitivity curves for stochastic gravitational- wave backgrounds that takes into account the increase in sensitivity that comes from integrating over frequency in addition to integrating over time. This method is valid for backgrounds that have a power-law spectrum in the analysis band. We call these graphs “power-law integrated curves.” For simplicity, we consider cross-correlation searches for unpolarized and isotropic stochastic back- grounds using two or more detectors. We apply our method to construct power-law integrated sen- sitivity curves for second-generation ground-based detectors such as Advanced LIGO, space-based detectors such as LISA and the Observer, and timing residuals from a timing array. The code used to produce these plots is available at https://dcc.ligo.org/LIGO-P1300115/public for researchers interested in constructing similar sensitivity curves.

I. INTRODUCTION

When discussing the feasibility of detecting gravita- tional waves using current or planned detectors, one often plots characteristic strain hc(f) curves of predicted sig- nals, and compares them to sensitivity curves for different detectors. The sensitivity curves are usually constructed by taking the ratio of the detector’s noise power spec- tral density Pn(f) to its sky- and polarization-averaged response to a (f), defining Sn(f) P (f)/ (f) and an effective characteristicR strain noise≡ n R amplitude hn(f) fSn(f). If the curve correspond- ing to a predicted≡ signal h (f) lies above the detector p c sensitivity curve hn(f) in some frequency band, then the signal has signal-to-noise ratio >1. An example of such a plot is shown in Fig. 1, which is taken from [1]. For stochastic gravitational waves, which are typically FIG. 1: Sensitivity curves for gravitational-wave observa- searched for by cross-correlating data from two or more tions and the predicted spectra of various gravitational-wave detectors, one often adjusts the height of a sensitivity sources, taken from [1]. curve to take into account the total observation time (e.g., T = 1 yr or 5 yr). For uncorrelated detector noise, the expected (power) signal-to-noise ratio of a cross- correlation search for a gravitational-wave background to-noise ratio ρ (see Eq. 21) also scales like √Nbins = for frequencies between f and f +δf scales like √Tδf. So ∆f/δf, where Nbins is the number of frequency bins the effective characteristic strain noise amplitude hn(f) of width δf in the total bandwidth ∆f. As we shall should be multiplied by a factor of 1/(Tδf)1/4. Also, seep below, the actual value of the proportionality con- instead of characteristic strain, one often plots the pre- stant depends on the spectral shape of the background dicted fractional energy density in gravitational waves and on the detector geometry (e.g., the separation and Ωgw(f) as a function of frequency, which is proportional relative orientation of the detectors), in addition to the 2 2 to f hc(f) (see Eq. 6). An example of such a plot is individual detector noise power spectral densities. Since shown in Fig. 2, which is taken from [2]. this improvement to the sensitivity is signal dependent, it But for stochastic gravitational waves, plots such as is not always folded into the detector sensitivity curves, Figs. 1 and 2 do not always tell the full story. Searches even though the improvement in sensitivity can be sig- for gravitational-wave backgrounds also benefit from the nificant.1 And when it is folded in, as in Fig 2, a single broadband nature of the signal. The integrated signal-

1 To be clear, integration over frequency is always carried out aElectronic address: ethrane@.caltech.edu in searches for stochastic gravitational-wave backgrounds, even bElectronic address: [email protected] though this is not always depicted in sensitivity curves. 2

−2 −4 10 10 LIGO S4 CMB & Matter BBN Spectra −6 10 Planck −4 LIGO S5 10 Pulsar −8 Cosmic Strings 10 Limit −6 gw (f)

Ω AdvLIGO 10 −10 Ω 10 LISA CMB Large Pre−Big−Bang −12 Angle −8 10 10

Inflation −14 10 −10 10 −18 −16 −14 −12 −10 −8 −6 −4 −2 0 2 4 6 8 10 1 2 3 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 Frequency (Hz) f (Hz)

FIG. 2: Plot showing strengths of predicted gravitational- FIG. 3: Ωgw(f) sensitivity curves from different stages in a po- wave backgrounds in terms of Ωgw(f) and the corresponding tential future Advanced LIGO Hanford-LIGO Livingston cor- sensitivity curves for different detectors, taken from [2]. Up- relation search for power-law gravitational-wave backgrounds. per limits from various measurements, e.g., S5 LIGO Hanford- The top black curve is the single-detector sensitivity curve, as- Livingston and pulsar timing, are shown as horizontal lines sumed to be the same for both H1 or L1. The red curve shows in the analysis band of each detector. The upper limits take the sensitivity of the H1L1 detector pair to a gravitational- into account integration over frequency, but only for a single wave background, where the spikes are due to zeros in the spectral index. Hanford-Livingston overlap reduction function (see left panel, Fig. 5). The green curve shows the improvement in sensitivity that comes from integration over an observation time of 1 year spectral index is assumed, making it difficult to compare for a frequency bin size of 0.25 Hz. The set of black lines are obtained by integrating over frequency for different power law published limits with arbitrary models. In other cases, indices, assuming a signal-to-noise ratio ρ = 1. Finally, the limits are given as a function of spectral index, but the blue power-law integrated sensitivity curve is the envelope of constrained quantity depends on an arbitrary reference the black lines. See Sec. III, Fig. 7 for more details. frequency; see Eq. 7. To illustrate the improvement in sensitivity that comes from integrating over frequency, consider the simple case of a white gravitational-wave background signal in white isotropic stochastic backgrounds using two or more de- uncorrelated detector noise. In this case, ρ increases by tectors. In Sec. III we present a graphical method for con- precisely √Nbins compared to the single bin analysis. For structing sensitivity curves for power-law backgrounds ground-based detectors like LIGO, typical values of ∆f based on the expected signal-to-noise ratio for the search, and δf are ∆f 100 Hz and δf 0.25 Hz, leading to and we apply our method to construct new power-law in- ≈ ≈ Nbins 400, and a corresponding improvement in ρ of tegrated sensitivity curves for correlation measurements about≈ 20. For colored spectra and non-trivial detector involving second-generation ground-based detectors such geometry the improvement will be less, but a factor of as Advanced LIGO, space-based detectors such as the Big 5-10 increase in ρ is not unrealistic. Bang Observer (BBO), and a . For ∼ In this paper, we propose a relatively simple way to completeness, we also construct a power-law integrated graphically represent this improvement in sensitivity for sensitivity curve for an autocorrelation measurement us- gravitational-wave backgrounds that have a power-law ing LISA. We conclude with a brief discussion in Sec. IV. frequency dependence in the sensitivity band of the de- tectors. An example of such a “power-law integrated sensitivity curve” is given in Fig. 3 for a correlation mea- surement between the Advanced LIGO detectors in Han- ford, WA and Livingston, LA. Details of the construction II. FORMALISM and interpretation of these curves will be given in Sec III, Fig. 7. We show this figure now for readers who might be anxious to get to the punchline. In this section, we summarize the fundamental prop- In Sec. II, we briefly review the fundamentals of cross- erties of a stochastic background and the correlated re- correlation searches for gravitational-wave backgrounds, sponse of a network of detectors to such a background. In defining an effective strain noise power spectral density order to keep track of the many different variables neces- Seff (f) for a network of detectors. For simplicity, we sary for this discussion, we have included Table I, which consider cross-correlation searches for unpolarized and summarizes key variables. 3

variable definition

hab(t, ~x) metric perturbation, Eq. 1

hA(f, kˆ) Fourier coefficients of metric perturbation, Eq. 1

Sh(f) strain power spectral density of a gravitational-wave background, Eq. 3

Ωgw(f) fractional energy density spectrum of a gravitational-wave background, Eq. 4

hc(f) characteristic strain for gravitational waves, Eq. 5 h(t) detector response to gravitational waves, Eq. 12 A ˆ RI (f, k) detector response to a sinusoidal plane gravitational wave, Eq. 12 h˜(f) Fourier transform of h(t), Eq. 13

ΓIJ (f) overlap reduction function for the correlated response to a gravitational-wave background, Eq. 15

I (f) detector response to a gravitational wave averaged over polarizations and directions on the sky, Eq. 17 R PhI (f) detector power spectral density due to gravitational waves, Eq. 18

PnI (f) detector power spectral density due to noise, Eq. 21

Seff (f) effective strain noise power spectral density for a detector network, Eq. 23

heff (f) effective characteristic strain noise amplitude for a detector network, Eq. 24

Sn(f) strain noise power spectral density for a single detector, Eq. 27

hn(f) characteristic strain noise amplitude for a single detector, hn(f) pfSn(f) ≡ TABLE I: Summary of select variables with references to key equations.

A. Statistical properties to close the universe [3]. (Throughout this paper we uti- lize single-sided power spectra.) The variable ρc denotes In transverse-traceless coordinates, the metric pertur- the critical energy density of the universe while dρgw de- bations hab(t, ~x) corresponding to a gravitational-wave notes the energy density between f and f + df. In terms background can be written as a linear superposition of of the characteristic strain sinusoidal plane gravitational waves with frequency f, hc(f) fSh(f) , (5) propagation direction kˆ, and polarization A: ≡ it follows that p hab(t, ~x)= 2π2 ∞ 2 2 Ωgw(f)= 2 f hc(f) . (6) 2 ˆ A ˆ i2πf(t−kˆ·~x/c) (1) 3H0 df d Ωkˆ hA(f, k)eab(k) e , −∞ S2 Z Z XA A ˆ B. Power-law backgrounds where eab(k) are the gravitational-wave polarization ten- sors and A =+, (see e.g., [3]). The Fourier components × In this paper, we will restrict our attention to hA(f, kˆ) are random fields whose expectation values de- fine the statistical properties of the background. With- gravitational-wave backgrounds that can be described by power-law spectra: out loss of generality we can assume h (f, kˆ) = 0. For h A i unpolarized and isotropic stochastic backgrounds, the f β Ω (f)=Ω , (7) quadratic expectation values have the form gw β f  ref  ˆ ∗ ′ ˆ′ hA(f, k)hA′ (f , k ) = where β is the spectral index and fref is a reference fre- h i −1 1 ′ ′ quency, typically set to 1 yr for pulsar-timing observa- ′ 2 ˆ ˆ δ(f f )δAA δ (k, k )Sh(f), (2) tions and 100 Hz for ground-based detectors. The choice 16π − of fref, however, is arbitrary and does not affect the de- where tectability of the signal. It follows trivially that the characteristic strain also 3H2 Ω (f) S (f)= 0 gw (3) has a power-law form: h 2π2 f 3 f α h (f)= A , (8) is the gravitational-wave power spectral density, and c α f  ref  1 dρgw where the amplitude Aα and spectral index α are related Ωgw(f)= (4) ρc d ln f to Ωβ and β via: 2 is the fractional contribution of the energy density in 2π 2 2 Ωβ = 2 fref Aα , β =2α +2 . (9) gravitational waves to the total energy density needed 3H0 4

For inflationary backgrounds relevant for cosmology, it to define a normalized overlap reduction function γIJ (f) is often assumed that such that for two identical, co-located and co-aligned de- tectors, γIJ (0) = 1. For identical interferometers with Ωgw(f) = const , (10) opening angle δ, for which β = 0 and α = 1. For a background arising 2 γIJ (f)=(5/ sin δ)ΓIJ (f) . (16) from binary coalescence, − For a single detector (i.e., I = J), we define Ω (f) f 2/3 , (11) gw ∝ (f) Γ (f), (17) for which β = 2/3 and α = 2/3. This power-law de- RI ≡ II − pendence is applicable to super-massive black-hole coa- which is the transfer function between gravitational-wave lescences targeted by pulsar timing observations as well strain power Sh(f) and detector response auto power as compact binary coalescences relevant for ground-based and space-based detectors. P (f)= (f)S (f) . (18) hI RI h

Note that I (f) is the antenna pattern of detector I C. Detector response averaged overR polarizations and directions on the sky. A plot of (f) normalized to unity for the strain re- RI The response h(t) of a detector to a passing gravita- sponse of an equal-arm Michelson interferometer is shown tional wave is the convolution of the metric perturbations in Fig. 4. ab hab(t, ~x) with the impulse response R (t, ~x): 0 ∞ 10 3 ab h(t) dτ d y R (τ, ~y)hab(t τ, ~x ~y) ≡ −∞ − − Z ∞ Z −1 2 A ˆ ˆ i2πf(t−kˆ·~x/c) 10 = df d Ωkˆ R (f, k)hA(f, k)e , −∞ Z Z A X (12) −2 (f) 10 II where ~x is the location of the measurement at time t. The γ function RA(f, kˆ) is the detector response to a sinusoidal

ˆ −3 plane-wave with frequency f, propagation direction k, 10 and polarization A. In the frequency domain, we have

2 A −i2πfkˆ·~x/c ˜ ˆ ˆ −4 h(f)= d Ωkˆ R (f, k)hA(f, k)e . (13) 10 −2 −1 0 1 Z A 10 10 10 10 X 2fL/c

D. Overlap reduction function FIG. 4: A plot of the transfer function I (f) = γII (f) normalized to unity for the strain responseR of an equal-arm Given two detectors, labeled by I and J, the expec- Michelson interferometer. The dips in the transfer function tation value of the cross-correlation of the detector re- occur around integer multiples of c/(2L), where L is the arm length of the interferometer. sponses h˜I (f) and h˜J (f) is

∗ ′ 1 ′ h˜ (f)h˜ (f ) = δ(f f )Γ (f)S (f) , (14) h I J i 2 − IJ h Detailed derivations and discussions of the overlap re- duction functions for ground-based laser interferometers, where space-based laser interferometers, and pulsar timing ar- rays can be found in [3–5], [6, 7], and [8, 9], respec- Γ (f) IJ ≡ tively. In Fig. 5 we plot the overlap reduction functions 1 2 A ˆ A∗ ˆ −i2πfkˆ·(~xI −~xJ )/c for the strain response of the LIGO Hanford-LIGO Liv- d Ωˆ RI (f, k)RJ (f, k)e 8π k ingston detector pair in the long-wavelength limit (valid Z A X (15) for frequencies below a few kHz), and the strain re- sponse of a pair of mini LISA-like Michelson interfer- is the overlap reduction function (see e.g., [4, 5] in the ometers in the hexagram configuration of the Big-Bang context of ground-based interferometers). Note that Observer (BBO), which is a proposed space-based mis- ΓIJ (f) is the transfer function between gravitational- sion, whose goal is the direct detection of the cosmo- wave strain power Sh(f) and detector response cross- logical gravitational-wave background [10–12]. The two power CIJ (f) = ΓIJ (f)Sh(f). It is often convenient Michelson interferometers for the BBO overlap reduction 5 function are located at opposite vertices of a hexagram with corresponding strain noise amplitude (‘Star of David’) and have arm lengths L = 5 107 m ◦ × and opening angles δ = 60 . heff (f) fSeff (f) . (24) ≡ In Fig. 6, we plot both the overlap reduction func- In terms of S (f), we havep tion and the Hellings and Downs curve [8] for the timing eff response of a pair of in a pulsar timing array. As- 2 1/2 Sh suming the pulsars are separated by an angle ψIJ on the ρ = 2Tδf N , (25) bins S2 sky, then to a very good approximation [9]:  eff  p p 1 1 where denotes an average3 over the total bandwidth ΓIJ (f)= ζIJ (19) h i (2πf)2 3 of the detectors, ∆f = Nbins δf. For the case of M iden- tical, co-located and co-aligned detectors, things simplify where further. First, 3 1 cos ψ 1 cos ψ ζ − IJ log − IJ IJ ≡2 2 2 2     (20) Seff (f)= Sn(f) , (26) 1 1 cos ψIJ 1 1 sM(M 1) − + + δ − − 4 2 2 2 IJ   where is the Hellings and Downs factor [8]. (The normalization is chosen so that for a single pulsar ζ = 1.) Sn(f) Pn(f)/ (f) (27) II ≡ R is the strain noise power spectral density in a single de- E. Signal-to-noise ratio tector. Second, S2 1/2 The expected (power) signal-to-noise ratio for a cross- ρ = Tδf N M(M 1) h . (28) bins − S2 correlation search for an unpolarized and isotropic  n  p p p stochastic background is given by [3]: Thus, we see that the expected signal-to-noise ratio ∞ 1/2 scales linearly with the number of detectors for M 1, Γ2 (f)S2(f) ≫ ρ = √2T df IJ h , (21) the square-root of the total observation time, and the P (f)P (f) Z0 nI nJ  square-root of the number of frequency bins. Note where T is the total (coincident) observation time and that √Tδf√Nbins = √T ∆f, which is the total time- PnI (f), PnJ (f) are the auto power spectral densities for frequency volume of the measurement. the noise in detectors I, J. This the total broadband signal-to-noise ratio, integrated over both time and fre- quency. It can be derived as the expected signal-to-noise III. POWER-LAW INTEGRATED CURVES ratio of a filtered cross-correlation of the output of two detectors, where the filter function is chosen so as to max- A. Construction imize the signal-to-noise ratio of the cross-correlation.2 For a network of detectors, this generalizes to The sensitivity curves that we propose are based on Eq. 22 for the expected signal-to-noise ratio ρ, applied ∞ 1/2 M M Γ2 (f)S2(f) √ IJ h to gravitational-wave backgrounds with power-law spec- ρ = 2T df , (22) tra. These “power-law integrated sensitivity curves” in- " 0 PnI (f)PnJ (f)# Z XI=1 J>IX clude the improvement in sensitivity that comes from the where M the number of individual detectors, and we have broadband nature of the signal, via the integration over assumed the same coincident observation time T for each frequency. The following construction is cast in terms detector. of Ωgw(f), but we note that power-law integrated curves The above expression for ρ suggests the following def- can also easily be constructed for hc(f) or Sh(f) using inition of an effective strain noise power spectral density Eqs. 3 and 5 to convert between the different quantities. for the detector network − 1. Begin with the detector noise power spectral den- M M 1/2 2 sities PnI (f), PnJ (f), and the overlap reduction ΓIJ (f) Seff (f) , (23) functions ΓIJ (f) for two or more detectors. Us- ≡ " PnI (f)PnJ (f)# XI=1 J>IX ing Eq. 23, first calculate the effective strain power spectral density Seff (f), and then convert it to en- ergy density units Ωeff (f) using Eq. 3.

2 The above expression for ρ assumes that the gravitational-wave background is weak compared to the instrumental noise in the sense that PhI (f) ≪ PnI (f) for all frequencies in the bandwidth 3 X 1 fmax X f df of the detectors. Explicitly, h i≡ ∆f Rfmin ( ) . 6

1.2

1

0.8

0.6 (f) γ 0.4

0.2

0

−0.2 −2 −1 0 1 10 10 10 10 f (Hz)

FIG. 5: Left panel: Normalized overlap reduction function for the LIGO detectors located in Hanford, WA and Livingston, LA. Right panel: Normalized overlap reduction function for two mini LISA-like Michelson interferometers located at opposite vertices of the BBO hexagram configuration.

16 10

14 10 ) 2 12 10 (f) (s Γ

10 10

8 10 −9 −8 −7 −6 10 10 10 10 f (Hz)

FIG. 6: Left panel: Overlap reduction function for a pair of pulsars, with ζIJ chosen to be 0.25. Right panel: Hellings and Downs function ζ(ψIJ ). Note that the overlap reduction function is a function of frequency for a fixed pair of pulsars, while the Hellings and Downs function is a function of the angle between two pulsars, and is independent of frequency.

β 2. Assume an observation time T , typically between Ωβ(f/fref) versus f. 1 and 10 yr. 5. The envelope of the Ωgw(f) power-law curves is the 3. For a set of power-law indices e.g., β = power-law integrated sensitivity curve for a corre- 8, 7, 7, 8 and some choice of reference fre- lation measurement using two or more detectors. quency{− − f· · ·, calculate} the value of the amplitude ref Interpretation: Any line (on a log-log plot) that is tan- Ω such that the integrated signal-to-noise ratio β gent to the power-law integrated sensitivity curve cor- has some fixed value, e.g., ρ = 1. Explicitly, responds to a gravitational-wave background power-law −1/2 spectrum with an integrated signal-to-noise ratio ρ = 1. fmax 2β ρ (f/fref) This means that if the curve for a predicted background Ωβ = df 2 , (29) √2T min Ω (f) lies everywhere below the sensitivity curve, then ρ< 1 for "Zf eff # such a background. On the other hand, if the curve for where [fmin,fmax] define the bandwidth ∆f of the a predicted power-law background with spectral index β detectors. Note that the choice of fref is arbitrary lies somewhere above the sensitivity curve, then it will be and will not affect the sensitivity curve. pred observed with an expected value of ρ = Ωβ /Ωβ > 1. pred 4. For each pair of values for β and Ωβ, plot Ωgw(f)= Graphically, Ωβ is the value of the predicted power-law 7 spectrum evaluated at fref , while Ωβ is the value of the (a) For the Advanced LIGO networks, we use the de- same power-law spectrum that is tangent to the sensitiv- sign detector noise power spectral density Pn(f) taken ity curve, also evaluated at fref . from [14] assumed to be the same for every detector in the network. We consider three networks: H1L1 (just the US aLIGO detectors), H1H2 (a hypothetical co-located pair B. Plots of aLIGO detectors), and H1L1V1K1 (the US aLIGO detectors plus detector pairs created with Virgo V1 and KAGRA K1).4 In reality, Virgo and KAGRA are ex- The calculation of a power-law integrated sensitivity pected to have different noise curves than aLIGO, but curve is demonstrated in the left-hand panel of Fig. 7 for we assume the same aLIGO noise for each detector in or- the Hanford-Livingston (H1L1) pair of Advanced LIGO der to show how the sensitivity curve changes by adding detectors. Following steps 1–5 above, we begin with the additional identical detectors to the network. Given this design detector noise power spectral density P (f) for n assumption, the effective strain power spectral density an Advanced LIGO detector [14] (which we assume to be can be written as the same for both H1 and L1), and divide by the H1L1 overlap reduction function to obtain the effective strain Seff (f)= Pn(f)/ eff (f) , (30) spectral density Seff (f) = Pn(f)/ΓH1L1(f) of the detec- R tor pair to a gravitational-wave background (see Eq. 23). where We then convert Seff (f) to an energy density Ωeff (f) via Eq. 3 to obtain the solid red curve. After integrating 1 yr M M 1/2 of coincident data, and assuming a frequency bin width (f)= Γ2 (f) (31) Reff IJ of 0.25 Hz, we obtain the solid green curve, which is lower "I=1 J>I # by a factor of 1/√2Tδf. Then assuming different spec- X X tral indices β, we integrate over frequency (see Eq. 29), is the sky- and polarization-averaged response of the net- setting ρ = 1 to determine the amplitude Ωβ of a power- work to a gravitational-wave background. A plot of the law background. This gives us the set of black lines for various overlap reduction functions γIJ (f) and eff (f) R each power law index β. The blue power-law integrated for the H1L1V1K1 network are given in Fig. 8. The re- curve is the envelope of these black lines. sulting power-law integrated sensitivity curves are shown The right-hand panel of Fig. 7 illustrates how to inter- in Fig. 9. pret a power-law integrated sensitivity curve. We replot (b) For the BBO sensitivity curve, the noise power the green and blue curves from the left-hand panel, which spectral density for the two Michelson interferometers is respectively represent the time-integrated and power-law taken to be integrated sensitivity of an Advanced LIGO H1L1 corre- 2 lation measurement to a gravitational-wave background. 4 2 (δa) Pn(f)= 2 (δx) + 4 , (32) Additionally, we plot two theoretical spectra of the form L " (2πf) # Ω (f) f 2/3, which is expected for a background due f gw f to compact∝ binary coalescences. The dark brown line where corresponds to a somewhat pessimistic scenario in which 2 − m Advanced LIGO, running at design sensitivity, would de- (δx)2 =2 10 34 , (33) tect 10 individual binary coalescences per × Hz ≈ 2 year of science data [13]. The light brown line repre- − m (fδa)2 =9 10 34 (34) sents a somewhat optimistic model in which Advanced × s4 Hz LIGO, running at design sensitivity, would detect 100 · individual binary neutron star coalescences per year≈ of are the position andf acceleration noise (see Table II from [11]) and L = 5 107 m is the arm length. Following science data [13]. (A binary-neutron-star detection rate × of 40yr−1 is considered a realistic rate for Advanced [12], we have included an extra factor of 4 multiplying the LIGO [15].) The light-brown curve intersects the blue first term in Eq. 32, which corresponds to high-frequency power-law integrated curve, indicating that the some- noise 4 times larger than shot noise alone. The overlap what optimistic model will induce a signal-to-noise ra- reduction function for the Michelson interferometers lo- tio ρ > 1. The dark brown curve falls below the blue cated at opposite vertices of the BBO hexagram is shown power-law integrated curve, indicating that the some- in the right panel of Fig. 5. The power-law integrated what pessimistic model will induce a signal-to-noise ratio curve for BBO is given in Fig. 10, top panel. ρ< 1. Note that neither curve intersects the green time- integrated sensitivity curve. In Figs. 9–11 we plot power-law integrated sensitiv- 4 ity curves for several upcoming or proposed experiments: We have taken the location and orientation of the KAGRA de- tector to be that of the TAMA 300-m interferometer in Tokyo, (a) networks of Advanced LIGO detectors, (b) BBO, (c) Japan. We have not included the planned LIGO India detec- LISA, and (d) a network of pulsars from a pulsar timing tor [16] in this network, as the precise LIGO-India site has not array. yet been decided upon. 8

0 −5 10 10

−2 −6 10 10

−4 −7 10 10 (f) (f) Ω Ω −6 −8 10 10

−8 −9 10 10

−10 −10 10 1 2 3 10 1 2 10 10 10 10 10 f (Hz) f (Hz)

FIG. 7: Left panel: Ωgw(f) sensitivity curves from different stages in a potential future Advanced LIGO H1L1 correlation search for power-law gravitational-wave backgrounds. The red line shows the effective strain spectral density Seff(f)= Pn(f)/ΓH1L1(f) of the H1L1 detector pair to a gravitational-wave background signal converted to energy density Ωeff (f) via Eq. 3. (The Pn(f) used in this calculation is the design detector noise power spectral density for an Advanced LIGO detector, assumed to be the same for both H1 and L1.) The spikes in the red curve are due to zeroes in the overlap reduction function ΓH1L1(f), which is shown in the left panel of Fig. 5. The green curve, Seff (f)/√2T δf, is obtained through the optimal combination of one year’s worth of data, assuming a frequency bin width of 0.25 Hz as is typical [2]. The vertical dashed orange line marks a typical Advanced LIGO reference frequency, fref = 100 Hz. The set of black lines are obtained by performing the integration in Eq. 29 for different power law indices β, requiring that ρ = 1 to determine Ωβ . Finally, the blue power-law integrated sensitivity curve is the envelope of the black lines. Right panel: a demonstration of how to interpret a power-law integrated curve. The thin green line and thick blue line are the same as in the left panel. The two dashed brown lines represent two different plausible signal models for gravitational-wave backgrounds arising from binary neutron star coalescence; see, e.g., [13]. In each case, 2/3 Ωgw(f) f ; however, the two curves differ by an order of magnitude in the overall normalization of Ωgw(f). The louder signal will∝ induce a signal-to-noise-ratio ρ > 1 with an Advanced LIGO H1L1 correlation measurement as it intersects the blue power-law integrated curve—even though it falls below the time-integrated green curve. The weaker signal will induce a signal-to-noise-ratio ρ< 1 with Advanced LIGO H1L1 as it is everywhere below the power-law integrated curve.

0.4 1 H1L1 H1K1 0.9 0.2 H1V1 L1K1 0.8 0 L1V1 K1V1 0.7 0.6 −0.2 (f)

(f) 0.5 eff γ

−0.4 R 0.4

−0.6 0.3 0.2 −0.8 0.1

−1 0 0 1 2 3 0 1 2 3 10 10 10 10 10 10 10 10 f (Hz) f (Hz)

FIG. 8: Left panel: Individual normalized overlap reduction functions for the six different detector pairs comprising the H1L1K1V1 network. Right panel: Sky- and polarization-averaged reponse of the H1L1V1K1 network to a gravitational-wave background.

(c) For LISA, the analysis is necessarily different since LISA constellation. This is because the two independent the standard cross-correlation technique used for multi- Michelson inteferometers that one can synthesize from ple detectors such as an Advanced LIGO network, BBO, the six links of the standard equilateral LISA configura- or a pulsar timing array is not possible for a single tion are rotated at 45◦ with respect one another, leading 9

−4 10 are the position and acceleration noise [6, 11] and L = iH1L1 95% 5 109 m is the arm length. The transfer function (f) is H1L1 taken× from Fig. 4 restricted to the LISA band, 10−R4 Hz < H1L1V1K1 f < 10−1 Hz. Using the above expression for ρ and H1H2 −6 following the same steps from the previous subsection 10 for the construction of a power-law integrated curve, we obtain the sensitivity curve for LISA given in Fig. 10, (f)

Ω middle panel. Note that the minimum value of Ω(f) shown in this −8 10 plot is about a factor of 10 times smaller than the value −13 of Ωgw(f) 2 10 reported in [20, 21]. Part of this difference≈ is due× to our use of ρ = 1 for the sensitiv- ity curve, while their value of Ωgw(f) corresponds to a −10 strong (several σ) detection having a Bayes factor 30. 10 1 2 3 10 10 10 The remaining factor can probably be attributed to≥ the f (Hz) marginalization over the instrumental noise and galactic foreground parameters in [20, 21], while Eq. 35 assumes FIG. 9: Different networks of advanced detectors assuming that we know these parameters perfectly. T = 1 yr of observation. We also include 95% CL limits from (d) For the pulsar timing array sensitivity curve, we initial LIGO for comparison [2]. consider a network of 20 pulsars taken from the Inter- national Pulsar Timing Network (IPTA) [22], which we to zero cross-correlation for an isotropic gravitational- assume have identical white timing noise power spectral wave background for frequencies below about c/2L = densities, 3 10−2 Hz [17]. It is possible, however, to construct a P (f)=2∆t σ2 , (39) combination× of the LISA data whose response to grav- n itational waves is highly suppressed at these frequen- where 1/∆t is the cadence of the measurements, taken to cies, and hence can be used as a real-time noise moni- be 20 yr−1, and σ is the root-mean-square timing noise, tor for LISA [18, 19]. It is also possible to exploit the taken to be 100 ns. We note that the pulsar timing net- differences between the transfer function and spectral work we envision may be somewhat optimistic as 100 ns shape of a gravitational-wave background and that due root-mean-square timing noise is ambitious. Also, we do to instrumental noise and/or an astrophysical foreground not include the loss of sensitivity arising from the fitting (e.g., from galactic white-dwarf binaries) to discriminate of each pulsar’s period P and spin-down rate P˙ . a gravitational-wave background from these other noise Since the timing noise power spectral densities are contributions [20, 21]. identical, it follows that For the ideal case of an autocorrelation measurement in − a single detector assuming perfect subtraction of instru- M M 1/2 mental noise and/or any unwanted astrophysical fore- 2 Seff (f)= Sn(f) ζIJ , (40) ground, Eq. 21 for the expected signal-to-noise ratio is "I=1 J>I # replaced by X X where ∞ 2 2 1/2 (f)Sh(f) ρ = √T df R , (35) 2 2 P 2(f) Sn(f)= Pn(f)/ (f)=12π f Pn(f) (41) Z0 n  R where (f) Γ(f) is the transfer function of the detec- and ζIJ are the Hellings and Downs factors for each pair R ≡ of pulsars in the array. For our choice of 20 pulsars, tor and Pn(f) is its noise power spectral density. (The √2 reduction in ρ compared to a cross-correlation analysis is M M due to the use of data from only one detector instead of ζ2 =4.74 , (42) two.) For standard LISA, IJ XI=1 J>IX 2 1 2 4(δa) which can thought of as the effective number of pulsar Pn(f)= 2 (δx) + 4 , (36) L " (2πf) # pairs for the network. Finally, we assume a total obser- f vation time T = 5 yr, which sets the lower frequency f where limit of Seff (f). Given these parameters, we expect the 2 pulsar timing array to be operating in the “intermediate − m (δx)2 =4 10 22 , (37) signal limit” [23]. We therefore utilize the scaling laws × Hz from Fig. 2 in Ref. [23] to adjust the power-law integrated 2 − m weak-signal (fδa)2 =9 10 30 (38) curves, since Eqs. 21, 22 for ρ are valid in the × s4 Hz limit and overestimate the expected signal-to-noise ratio · f 10

−8 10

−10 10

−12 10 (f) Ω −14 10

−16 10

−18 10 −2 −1 0 1 10 10 10 10 f (Hz)

−6 10

−8 10

−10 (f) 10 Ω

−12 10

−14 10 −4 −3 −2 −1 10 10 10 10 f (Hz)

−7 10

−8 10

−9 (f) 10 Ω

−10 10

−11 10 −9 −8 −7 10 10 10 f (Hz)

FIG. 10: One-sigma, power-law integrated sensitivity curves. The dashed purple curves show the effective strain spectral density Seff(f) (Sn(f) for LISA, middle panel) converted to fractional energy density units (see Eqs. 23, 27, 3). Top panel: BBO assuming T = 1 yr of observation. The spike at 2.5 Hz is due to a zero in the BBO overlap reduction function. Middle panel: LISA autocorrelation measurement assuming ≈T = 1 yr of observation and perfect subtraction of instrumental noise and/or any unwanted astrophysical foreground. Bottom panel: A pulsar timing array consisting of 20 pulsars, 100 ns timing noise, T = 5 yr of observation, and a cadence of 20 yr−1. 11 by a factor of 5 for an observation of T = 5 yr. The LIGO-P1300115/public for public download. Hopefully, power-law integrated≈ curve for IPTA is given in Fig. 10, this will allow other researchers to easily construct similar bottom panel. sensitivity curves. Required inputs are the noise power It is interesting to note that the power-law integrated spectral density PnI (f) for each detector in the network curves for Advanced LIGO and BBO are relatively round and the overlap reduction function ΓIJ (f) for each detec- in shape, whereas the pulsar timing curve is pointy. (The tor pair. Common default files are available for download steep Ω(f) f 5 spectrum can be understood as follows: with the plotting code. the transfer∝ function (f) contributes a factor of f 2 while the conversion from powerR to energy density contributes Although the above discussion has focused on com- an additional factor of f 3.) This reflects the fact that paring predicted strengths of gravitational-wave back- the sensitivity of pulsar timing measurements is mostly grounds to sensitivity curves for current or planned de- determined by a small band of the lowest frequencies in tectors, one can also present measured upper limits for the observing band regardless of the spectral shape of the power-law backgrounds in a similar way. That is, instead signal. of plotting the upper limits for Ωβ (for fixed fref) as a Finally, Fig. 11 shows the power-law integrated sensi- function of the spectral index β as in [2, 24, 25], one can tivity curves for the different detectors on a single plot plot the envelope of upper-limit power-law curves as a spanning a wide range of frequencies. function of frequency. This would better illustrate the frequency dependence of the upper limits in the observ- ing band of the detectors. IV. DISCUSSION Acknowledgments

We have presented a graphical representation of de- tector sensitivity curves for power-law gravitational-wave We thank Vuk Mandic and Nelson Christensen for backgrounds that takes into account the enhancement in helpful comments regarding an earlier draft of the paper. sensitivity that comes from integrating over frequency JDR would also like to thank Paul Demorest, Justin El- in addition to integrating over time. We applied this lis, Shane Larson, Alberto Sesana, and Alberto Vecchio method to construct new power-law integrated sensitivity for discussions related to PTA sensitivity curves. ET is a curves for cross-correlation searches involving advanced member of the LIGO Laboratory, supported by funding ground-based detectors, BBO, and a network of pulsars from United States National Science Foundation. LIGO from a pulsar timing array. We also constructed a power- was constructed by the California Institute of Technology law integrated sensitivity curve for an autocorrelation and Massachusetts Institute of Technology with funding measurement using LISA. The new curves paint a more from the National Science Foundation and operates un- accurate picture of the expected sensitivity of upcom- der cooperative agreement PHY-0757058. JDR acknowl- ing observations. The code that we used to produce edges support from NSF Awards PHY-1205585, PHY- the new curves is available at https://dcc.ligo.org/ 0855371, and CREST HRD-1242090.

[1] G. Hobbs, in High-Energy Emission from Pulsars and [13] C. Wu, V. Mandic, and T. Regimbau, Phys. Rev. D 85, their Systems, edited by D. F. Torres and N. Rea 104024 (2012). (Springer Berlin Heidelberg, 2011), Astrophysics and [14] D. Shoemaker, Advanced ligo anticipated sensitiv- Space Science Proceedings, p. 229. ity curves (2010), URL https://dcc.ligo.org/ [2] B. Abbott et al., Nature 460, 990 (2009). LIGO-T0900288/public. [3] B. Allen and J. D. Romano, Phys. Rev. D 59, 102001 [15] J. Abadie et al., Class. Quantum Grav. 27, 173001 (1999). (2010). [4] N. Christensen, Physical Review D 46, 5250 (1992). [16] B. Iyer et al., LIGO-India Tech. Rep. (2011), URL [5] E.´ E.´ Flanagan, Physical Review D 48, 2389 (1993). https://dcc.ligo.org/cgi-bin/DocDB/ShowDocument? [6] N. J. Cornish and S. L. Larson, Class. Quantum Grav. docid=75988. 18, 3473 (2001). [17] C. Cutler, Phys. Rev. D 57, 7089 (1998). [7] L. S. Finn, S. L. Larson, and J. D. Romano, Phys. Rev. [18] M. Tinto, J. W. Armstrong, and F. B. Estabrook, Phys. D 79, 062003 (2009). Rev. D 63, 021101 (2000). [8] R. W. Hellings and G. S. Downs, Astrophys. J. Lett. 265, [19] C. J. Hogan and P. L. Bender, Phys. Rev. D 64, 062002 L39 (1983). (2001). [9] M. Anholm, S. Ballmer, J. D. E. Creighton, L. R. Price, [20] M. R. Adams and N. J. Cornish, Phys. Rev. D 82, 022002 and X. Siemens, Phys. Rev. D 79, 084030 (2009). (2010). [10] S. Phinney et al., NASA Mission Concept Study (2004). [21] M. R. Adams and N. J. Cornish (2013), http://arxiv. [11] J. Crowder and N. J. Cornish, Phys. Rev. D 72, 083005 org/abs/1307.4116v2. (2005). [22] G. Hobbs et al., Class. Quantum Grav. 27, 084013 [12] C. Cutler and J. Harms, Phys. Rev. D 73, 042001 (2006). (2010). 12

−6 10 PTA −8 LISA 10 BBO LIGO H1L1 −10 10

−12 (f) 10 Ω

−14 10

−16 10

−18

10 −18 −16 −14 −12 −10 −8 −6 −4 −2 0 2 4 10 10 10 10 10 10 10 10 10 10 10 10 f (Hz)

FIG. 11: One-sigma, power-law integrated sensitivity curves for the different detectors considered in this paper, plotted on the same graph. The Advanced LIGO H1L1, BBO, and pulsar timing sensitivity curves correspond to correlation measurements using two or more detectors. The LISA sensitivity curve corresponds to an autocorrelation measurement in a single detector assuming perfect subtraction of instrumental noise and/or any unwanted astrophysical foreground.

[23] X. Siemens, J. Ellis, F. Jenet, and J. D. Romano (2013), [25] V. Mandic, E. Thrane, S. Giampanis, and T. Regimbau, http://arxiv.org/abs/1305.3196. Phys. Rev. Lett. 109, 171102 (2012). [24] J. Abadie et al., Phys. Rev. D 85, 122001 (2012).