1:
2:
3: \documentstyle[12pt,aps,epsf]{revtex}
4: \begin{document}
5: \tighten
6: \draft
7: \title{\bf{Relic Gravitational Waves and Their Detection}}
8:
9: \author{L. P. Grishchuk\thanks{e-mail: grishchuk@astro.cf.ac.uk}}
10: \address{Department of Physics and Astronomy, Cardiff University,
11: Cardiff CF2 3YB, United Kingdom \\ and \\ Sternberg Astronomical
12: Institute, Moscow University, Moscow 119899, Russia}
13:
14: \maketitle
15: \vspace{2cm}
16:
17: \begin{abstract}
18: The range of expected amplitudes and spectral slopes
19: of relic (squeezed) gravitational waves, predicted by theory and
20: partially supported by observations, is within the reach of
21: sensitive gravity-wave detectors. In the most favorable case, the detection of
22: relic gravitational waves can be achieved by the cross-correlation of outputs
23: of the initial laser interferometers
24: in LIGO, VIRGO, GEO600. In the more realistic case, the sensitivity of
25: advanced ground-based and space-based laser interferometers will be needed.
26: The specific statistical signature of relic gravitational waves, associated
27: with the phenomenon of squeezing, is a potential reserve for further
28: improvement of the signal to noise ratio.
29: \end{abstract}
30:
31:
32: %\end{titlepage}
33:
34: \vspace{1.5cm}
35:
36:
37:
38:
39: \section{Introduction}
40:
41: It is appropriate and timely to discuss the detection of relic
42: gravitatational waves at the experimental meeting like this one. We are
43: in the situation when the advanced laser interferometers, currently under
44: construction or in a design phase, can make the dream of detecting relic
45: gravitons a reality. The detection of relic gravitational waves is
46: the only way to learn about the evolution of the very early Universe,
47: up to the limits of Planck era and Big Bang.
48: \par
49: The existence of relic gravitational waves is a consequence of quite
50: general assumptions. Essentially, we rely only on the validity of general
51: relativity and basic principles of quantum field theory. The strong
52: variable gravitational field of the early Universe amplifies the inevitable
53: zero-point quantum oscillations of the gravitational waves and produces
54: a stochastic background of relic gravitational waves measurable
55: today \cite{g1}.
56: It is important to appreciate the fundamental and unavoidable nature of this
57: mechanism. Other physical processes can also generate stochastic backgrounds
58: of gravitational waves. But those processes either involve many additional
59: hypotheses, which may turn out to be not true, or produce a gravitational wave
60: background (like the one from binary stars in the Galaxy) which should
61: be treated as an unwanted noise rather than a useful and interesting signal.
62: The scientific importance of detecting relic gravitational waves has been
63: stressed on several occasions (see, for example, \cite{t1}--\cite{sch}).
64: \par
65: The central notion in the theory of relic gravitons is the phenomenon
66: of superadiabatic (parametric) amplification. The roots of this phenomenon
67: are known in classical physics, and we will remind its basic features. As
68: every wave-like process, gravitational waves use the concept of a harmonic
69: oscillator. The fundamental equation for a free harmonic oscillator is
70: \begin{equation}
71: \label{1}
72: \ddot{q} + \omega^2 q = 0,
73: \end{equation}
74: where $q$ can be a displacement of a mechanical pendulum or a
75: time-dependent amplitude of a mode of the physical field. The energy of the
76: oscillator can be changed by an acting force or, alternatively, by a
77: parametric influence, that is,
78: when a parameter of the oscillator, for instance the length of a pendulum,
79: is being changed. In the first case, the fundamental equation takes the
80: form
81: \begin{equation}
82: \label{2}
83: \ddot{q} + \omega^2 q = f(t),
84: \end{equation}
85: whereas in the second case Eq. (\ref{1}) takes the form
86: \begin{equation}
87: \label{3}
88: \ddot{q} + \omega^2 (t) q = 0.
89: \end{equation}
90: Equations (\ref{2}) and (\ref{3}) are profoundly different, both,
91: mathematically and physically.
92: \par
93: Let us concentrate on the parametric influence. We consider a pendulum
94: of length $L$ oscillating in a constant gravitational field $g$.
95: The unperturbed pendulum
96: oscillates with the constant frequency $\omega = \sqrt{g/L}$.
97: $Fig. 1a$ illustrates the
98: variation of the length of the pendulum $L(t)$ by an external agent,
99: shown by alternating arrows. Since $L(t)$ varies, the frequency
100: of the oscillator does also vary: $\omega(t) = \sqrt{g/L(t)}$.
101: The variation $L(t)$ does not
102: need to be periodic, but cannot be too much slow (adiabatic) if the result
103: of the process is going to be significant. Otherwise, in the adiabatic
104: regime of slow variations, the energy of the oscillator $E$ and its
105: frequency $\omega$ do change slowly, but $E /\omega$ remains constant, so
106: one can say that the ``number of quanta"
107: $E/ \hbar \omega$ in the oscillator remains fixed.
108: In other words, for the creation of new ``particles - excitations",
109: the characteristic time of the variation should be
110: comparable with the period of the oscillator and the adiabatic behaviour
111: should be violated. After some duration of the appropriate parametric influence,
112: the pendulum will oscillate at the original frequency, but will have a
113: significantly larger, than before, amplitude and energy.
114: This is shown in $Fig. 1b$. Obviously, the energy of the oscillator has
115: been increased at the expense of the external agent (pump field). For
116: simplicity, we have considered a familiar case, when the
117: length of the pendulum varies, while the gravitational acceleration $g$
118: remains constant. Variation of $g$ would represent a gravitational parametric
119: influence and would be even in a closer analogy with what we study below.
120: \par
121: A classical oscillator must have a non-zero initial amplitude for the
122: amplification mechanism to work. Otherwise, if
123: the initial amplitude is zero, the final amplitude will also be zero.
124: Indeed, imagine the pendulum strictly at rest, hanging stright down.
125: Whatever the variation of its length is, it will not make the pendulum
126: to oscillate and gain energy. In contrast,
127: a quantum oscillator does not need to be excited from the very beginning.
128: The oscillator can be
129: initially in its vacuum quantum-mechanical state. The inevitable zero-point
130: quantum oscillations are associated with the vacuum state energy
131: $\frac{1}{2} \hbar \omega$. One can imagine a pendulum hanging stright down,
132: but fluctuating with a tiny amplitude determined by the ``half of the
133: quantum in the mode". In the classical picture, it is this tiny amplitude
134: of quantum-mechanical origin that is being parametrically amplified.
135:
136: %Fig. 1
137: \begin{figure}
138: \epsfxsize=0.8\textwidth
139: \centerline{\epsfbox{oscil.fig1}}
140: \end{figure}
141:
142: \par
143: The Schrodinger evolution of a quantum oscillator depends crucially
144: on whether the oscillator is being excited parametrically or by a force.
145: Consider the phase diagram $(q, p)$, where $q$ is the
146: displacement and $p$ is the conjugate momentum. The vacuum state is
147: described by the circle in the center (see $Fig. 2$). The mean values of $q$
148: and $p$ are zeros, but their variances (zero-point quantum fluctuations)
149: are not zeros and are equal to
150: each other. Their numerical values are represented by the circle in the
151: center. Under the action of a force, the vacuum state evolves into a coherent
152: state. The mean values of $p$ and $q$ have increased, but the variances are
153: still equal and are described by the circle of the same size as for the
154: vacuum state. On the other hand,
155: under a parametric influence, the vacuum state evolves into a squeezed
156: vacuum state. [For a recent review of squeezed states see, for example,
157: \cite{kn} and references there.] Its variances for the
158: conjugate variables $q$ and $p$
159: are significantly unequal and are described by an ellipse.
160: As a function of time, the ellipse rotates with respect to the origin
161: of the $(q, p)$ diagram, and the numerical values of the variances
162: oscillate too. The mean numbers of quanta in the
163: two states, one of which is coherent and another is squeezed vacuum,
164: can be equal (similar to the coherent and squeezed states shown in $Fig. 2$)
165: but the statistical properties of these states are significantly different.
166: Among other things, the variance
167: of the phase of the oscillator in a squeezed vacuum state is very
168: small (squeezed). Graphically,
169: this is reflected in the fact that the ellipse is very thin, so that that
170: the uncertainty in the angle between the horizontal axis and the orientation
171: of the ellipse is very small. This highly elongated ellipse can be
172: regarded as a portarait of the gravitational wave quantum state that is being
173: inevitably generated by parametric amplification, and which we will be
174: dealing with below.
175:
176:
177: %Fig. 2
178: \begin{figure}
179: \epsfxsize=0.8\textwidth
180: \centerline{\epsfbox{squeeze.fig2}}
181: \end{figure}
182:
183:
184: A wave-field is not a single oscillator, it depends on spatial coordinates
185: and time, and may have several independent components (polarization
186: states). However, the field can be decomposed into a set of
187: spatial Fourier harmonics. In this way we represent the gravitational
188: wave field as a collection
189: of many modes, many oscillators. Because of the nonlinear character of
190: the Einstein equations, each of these oscillators is coupled to the variable
191: gravitational field of the surrounding Universe. For sufficiently
192: short gravitational waves of experimental interest, this coupling was
193: especially effective in the early Universe, when the condition of
194: the adiabatic behaviour of the oscillator was violated. It is this homogeneous
195: and isotropic gravitational field of all the matter in the early
196: Universe that played the role of the external agent - pump field.
197: The variable pump field acts parametrically on
198: the gravity-wave oscillators and drives them into multiparticle
199: states. Concretely, the initial vacuum state of each pair of waves with
200: oppositely directed momenta evolves into a highly correlated state
201: known as the two-mode squeezed vacuum state \cite{gs}, \cite{g2}.
202: The strength and duration of the effective coupling depends
203: on the oscillator's frequency. They all start in the vacuum state
204: but get excited to various amounts. As a result, a broad
205: spectrum of relic gravitational waves is being formed. This spectrum
206: is accessible to our observations today.
207: \par
208: Let us formulate the problem in more detail.
209:
210:
211: \section{Cosmological Gravitational Waves}
212:
213: In the framework of general relativity, a homogeneous isotropic gravitational
214: field is decribed by the line element
215: \begin{equation}
216: \label{4}
217: {\rm d}s^2 = c^2{\rm d}t^2 - a^2(t) \delta_{ij} {\rm d}x^i{\rm d}x^j =
218: a^2({\eta})[{\rm d}\eta^2 - \delta_{ij} {\rm d}x^i{\rm d}x^j].
219: \end{equation}
220: In cosmology, the function $a(t)$ (or $a(\eta)$) is called scale factor.
221: In our discussion, it will represent gravitational pump field.
222: \par
223: Cosmological gravitational waves are small corrections $h_{ij}$ to the metric
224: tensor. They are defined by the expression
225: \begin{equation}
226: \label{5}
227: {\rm d}s^2 = a^2({\eta})[{\rm d}\eta^2 - (\delta_{ij} + h_{ij})
228: {\rm d}x^i{\rm d}x^j].
229: \end{equation}
230: The functions $h_{ij} (\eta ,{\bf x})$ can be expanded over spatial
231: Fourier harmonics $e^{i{\bf nx}}$ and $e^{-i{\bf nx}}$, where
232: ${\bf n}$ is a constant wave vector. In this way, we reduce the dynamical
233: problem to the evolution of time-dependent amplitudes for each
234: mode ${\bf n}$.
235: Among six functions $h_{ij}$ there are only two independent (polarization)
236: components. This decomposition can be made, both, for real and for quantized
237: field $h_{ij}$. In the quantum version, the functions $h_{ij}$ are
238: treated as quantum-mechanical operators. We will use the Heisenberg
239: picture in which the time evolution is carried out by the operators while
240: the quantum state is fixed. This picture is fully equivalent to the
241: Schrodinger picture, discussed in the Introduction, in which the vacuum
242: state evolves into a squeezed vacuum state while the operators are time
243: independent.
244: \par
245: The Heisenberg operator for the quantized real field $h_{ij}$ can
246: be written as
247: \begin{eqnarray}
248: \label{6}
249: h_{ij} (\eta ,{\bf x})
250: = {C\over (2\pi )^{3/2}} \int_{-\infty}^\infty d^3{\bf n}
251: \sum_{s=1}^2~{\stackrel{s}{p}}_{ij} ({\bf n})
252: {1\over \sqrt{2n}}
253: \left[ {\stackrel{s}{h}}_n (\eta ) e^{i{\bf nx}}~
254: {\stackrel{s}{c}}_{\bf n}
255: +{\stackrel{s}{h}}_n^{\ast}(\eta ) e^{-i{\bf nx}}~
256: {\stackrel{s}{c}}_{\bf n}^{\dag} \right],
257: \end{eqnarray}
258: where $C$ is a constant which will be discussed later.
259: The creation and annihilation operators satisfy the conditions
260: $[{\stackrel{s'}{c}}_{\bf n},~{\stackrel{s}{c}}_{{\bf m}}^{\dag}]=
261: \delta_{s's}\delta^3({\bf n}-{\bf m})$,
262: ${\stackrel{s}{c}}_{\bf n}|0\rangle =0$, where $|0\rangle$
263: (for each ${\bf n}$ and $s$) is the fixed
264: initial vacuum state discussed below.
265: The wave number $n$ is related with the wave vector ${\bf n}$ by
266: $n = (\delta_{ij}n^in^j)^{1/2}$.
267: The two polarization tensors ${\stackrel{s}{p}}_{ij}({\bf n})$ $(s = 1, 2)$
268: obey the conditions
269: \[
270: {\stackrel{s}{p}}_{ij}n^j = 0, ~~
271: {\stackrel{s}{p}}_{ij}\delta^{ij} = 0, ~~
272: {\stackrel{s'}{p}}_{ij}
273: {\stackrel{s}{p}}~^{ij} = 2\delta_{ss'}, ~~
274: {\stackrel{s}{p}}_{ij}(-{\bf n}) = {\stackrel{s}{p}}_{ij}({\bf n}).
275: \]
276: The time evolution, one and the same for all ${\bf n}$ belonging to a
277: given $n$, is represented by the complex time-dependent function
278: ${\stackrel{s}{h}}_n(\eta )$. This evolution is dictated by the
279: Einstein equations. The nonlinear nature of the Einstein equations
280: leads to the coupling of
281: ${\stackrel{s}{h}}_n(\eta )$ with the pump field $a(\eta)$.
282: For every wave number $n$
283: and each polarization component $s$,
284: the functions ${\stackrel{s}{h}}_n(\eta )$ have the form
285: \begin{equation}
286: \label{7}
287: {\stackrel{s}{h}}_n(\eta ) = {1\over a(\eta )}
288: [{\stackrel{s}{u}}_n(\eta ) + {\stackrel{s}{v}}_n^{\ast} (\eta )],
289: \end{equation}
290: where ${\stackrel{s}{u}}_n(\eta )$ and ${\stackrel{s}{v}}_n(\eta )$
291: can be expressed in terms of the three
292: real functions (the polarization index $s$ is omitted):
293: $r_n$ - squeeze parameter,
294: $\phi_n$ - squeeze angle, $\theta_n$ - rotation angle,
295: \begin{equation}
296: \label{8}
297: u_n = e^{i{\theta}_n} \cosh~{r}_n, \qquad
298: v_n = e^{-i({\theta}_n - 2{\phi}_n )} \sinh~{r}_n.
299: \end{equation}
300: The dynamical equations for $u_n(\eta)$ and $v_n(\eta)$
301: \begin{equation}
302: \label{9}
303: i\frac{{\rm d} u_n}{{\rm d}\eta} = n u_n + i\frac{a'}{a} v_n^{*}, \qquad
304: i\frac{{\rm d} v_n}{{\rm d}\eta} = n v_n + i\frac{a'}{a} u_n^{*}
305: \end{equation}
306: lead to the dynamical equations governing
307: the functions $r_n(\eta)$, $\phi_n(\eta)$, $\theta_n(\eta)$ \cite{g2}:
308: \begin{equation}
309: \label{10}
310: r_n^{\prime} = \frac{a^{\prime}}{a} \cos{2{\phi}_n}, \quad
311: \phi_n^{\prime} = -n - \frac{a^{\prime}}{a} \sin{2{\phi}_n}\coth~2{r}_n, \quad
312: \theta_n^{\prime} = -n - \frac{a^{\prime}}{a} \sin{2{\phi}_n}\tanh~{r}_n,
313: \end{equation}
314: where $^{\prime} = {\rm d}/{\rm d}\eta$, and the evolution begins
315: from $r_n = 0$. This value of $r_n$
316: characterizes the initial vacuum state $|0\rangle$ which is defined
317: long before the interaction with the pump field
318: became effective, that is, long before the coupling term
319: $a^{\prime}/a$ became comparable with $n$.
320: The constant $C$ should be taken as
321: $C=\sqrt{16\pi}~l_{Pl}$ where $l_{Pl}=(G\hbar /c^3)^{1/2}$ is
322: the Planck length. This particular value of the constant $C$
323: guarantees the correct quantum normalization of the field: energy
324: $\frac{1}{2} \hbar \omega$ per each mode in the initial vacuum state.
325: The dynamical equations and their
326: solutions are identical for both polarization components $s$.
327: \par
328: Equations (\ref{9})
329: can be translated into the more familiar form of the second-order
330: differential equation for the function
331: ${\stackrel{s}{\mu}}_n(\eta ) \equiv
332: {\stackrel{s}{u}}_n(\eta ) + {\stackrel{s}{v}}_n^{\ast} (\eta )
333: \equiv a(\eta) {\stackrel{s}{h}}_n(\eta )$ \cite{g1}:
334: \begin{equation}
335: \label{11}
336: \mu_{n}^{\prime\prime} + \mu_{n} \left[n^2 -
337: \frac{a^{\prime\prime}}{a}\right] = 0.
338: \end{equation}
339: \par
340: Clearly, this is the equation for a parametrically disturbed oscillator
341: (compare with Eq. (\ref{3})). In absence of the gravitational parametric
342: influence represented by the term
343: $a^{\prime\prime}/a$, the frequency of the oscillator defined in terms
344: of $\eta$-time would be a constant: $n$. Whenever the term
345: $a^{\prime\prime}/a$ can be neglected, the general solution
346: to Eq. (\ref{11}) has the usual oscillatory form
347: \begin{equation}
348: \label{12}
349: \mu_n (\eta) = A_n e^{-in\eta} + B_n e^{in\eta},
350: \end{equation}
351: where the constants $A_n$, $B_n$ are determined by the initial conditions.
352: On the other hand, whenever the term $a^{\prime\prime}/a$ is dominant,
353: the general solution to Eq. (\ref{11}) has the form
354: \begin{equation}
355: \label{13}
356: \mu_n (\eta) = C_n a + D_n a \int^{\eta}\frac{{\rm d} \eta}{a^2}.
357: \end{equation}
358: In fact, this approximate solution is valid as long as $n$ is small in
359: comparison with $|a^{\prime}/a|$. This is more clearly seen from the
360: equivalent form of Eq. (\ref{11}) written in terms of the
361: function $h_n (\eta)$ \cite{ll}:
362: \begin{equation}
363: \label{14}
364: h_n^{\prime\prime} + 2\frac{a^{\prime}}{a} h_n^{\prime} + n^2 h_n = 0.
365: \end{equation}
366: For growing functions $a(\eta)$, that is, in expanding universes, the
367: second term in Eq.(\ref{13}) is usually smaller than the first one (see below),
368: so that, as long as $n \ll a^{\prime}/a$, the dominant solution is the growing
369: function $\mu_n (\eta) = C_n a(\eta)$, and
370: \begin{equation}
371: \label{d}
372: h_n = const.
373: \end{equation}
374: \par
375: Equation (\ref{11}) can be also looked at as a kind of the Schrodinger
376: equation for a particle moving in presence of the effective
377: potential $U(\eta) = a^{\prime\prime}/a$. In the situations that
378: are normally considered, the potential
379: $U(\eta)$ has a bell-like shape and forms a barrier (see $Fig. 3$).
380: When a given mode $n$ is outside the barrier, its amplitude $h_n$
381: is adiabatically decreasing with time:
382: $h_n \propto \frac{e^{{\pm}in\eta}}{a(\eta)}$.
383: This is shown in $Fig. 3$ by oscillating lines with decreasing amplitudes
384: of oscillations. The modes with sufficiently high frequencies do not
385: interact with the barrier, they stay above the barrier. Their amplitudes
386: $h_n$ behave adiabatically all the time. For
387: these high-frequency modes, the initial vacuum state (in the Schrodinger
388: picture) remains the vacuum forever. On the other hand, the modes that
389: interact with the barrier are subject to the superadiabatic amplification.
390: Under the barrier and as long as
391: $n < a'/a$, the function $h_n$ stays constant instead of the adiabatic
392: decrease. For these modes, the initial vacuum state evolves into a
393: squeezed vacuum state.
394:
395:
396: %Fig. 3
397: \begin{figure}
398: \epsfxsize=0.8\textwidth
399: \centerline{\epsfbox{barrier.fig3}}
400: \end{figure}
401:
402:
403:
404:
405: After having formulated the initial conditions, the present day behaviour
406: of $r_n$, $\phi_n$, $\theta_n$ (or, equivalently, the present day behaviour
407: of $h_n$) is
408: essentially all we need to find. The mean number of particles
409: in a two-mode squeezed state is $2\sinh^2{r_n}$ for each $s$. This number
410: determines the
411: mean square amplitude of the gravitational wave field. The time behaviour of
412: the squeeze angle $\phi_n$ determines the time dependence of the correlation
413: functions of the field. The amplification (that is, the growth of $r_n$)
414: governed
415: by Eq. (\ref{10}) is different for different wave numbers $n$. Therefore,
416: the present day results depend on the present day frequency $\nu$
417: ($\nu = {cn}/{2 \pi a}$) measured in $Hz$.
418: \par
419: In cosmology, the function $H \equiv \dot a/a \equiv c a^{\prime}/a^2$
420: is the time-dependent Hubble parameter. The function $l \equiv c/H$
421: is the time-dependent Hubble radius. The time-dependent wavelength of the
422: mode $n$ is $\lambda = 2 \pi a/n$. The wavelength $\lambda$ has this universal
423: definition in all regimes. In contrast, the $\nu$ defined
424: as $\nu = {cn}/{2 \pi a}$ has the usual meaning of a frequency of an
425: oscillating process only in the short-wavelength (high-frequency) regime of
426: the mode $n$, that is, in the regime where $\lambda \ll l$.
427: As we have seen above, the qualitative
428: behaviour of solutions to Eqs. (\ref{11}), (\ref{14}) depends
429: crucially on the comparative
430: values of $n$ and $a'/a$, or, in other words, on the comparative
431: values of $\lambda(\eta)$ and $l(\eta)$. This relationship is also
432: crucial for solutions to Eq. (\ref{10}) as we shall see now.
433: \par
434: In the short-wavelength regime, that is, during intervals of time
435: when the wavelength $\lambda(\eta)$ is shorter than the Hubble
436: radius $l(\eta) = a^2/a^{\prime}$, the term $n$ in (\ref{10}) is dominant.
437: The functions $\phi_n(\eta)$ and
438: $\theta_n(\eta)$ are $\phi_n = -n(\eta + \eta_n)$, $\theta_n = \phi_n$
439: where $\eta_n$ is a constant.
440: The factor $\cos 2\phi_n$ is a quickly oscillating
441: function of time, so the squeeze parameter $r_n$ stays practically constant.
442: This is the adiabatic regime for a given mode.
443: \par
444: In the opposite, long-wavelength regime, the term $n$ can be
445: neglected.
446: The function $\phi_n$ is $\tan \phi_n(\eta) \approx const/a^2(\eta)$,
447: and the squeeze angle
448: quickly approaches one of the two values: $\phi_n = 0$ or $\phi_n = \pi$
449: (analog of ``phase bifurcation" \cite{w}).
450: The squeeze parameter $r_n(\eta)$ grows with time according to
451: \begin{equation}
452: \label{15}
453: r_n(\eta) \approx ln \frac{a(\eta)}{a_*}~,
454: \end{equation}
455: where $a_*$ is the value of $a(\eta)$ at $\eta_*$, when the
456: long-wavelength regime, for a given $n$, begins. The final amount of $r_n$ is
457: \begin{equation}
458: \label{16}
459: r_n \approx ln \frac{a_{**}}{a_*}~,
460: \end{equation}
461: where $a_{**}$ is the value of $a(\eta)$ at $\eta_{**}$, when the
462: long-wavelength regime and
463: amplification come to the end. It is important to emphasize that it is not
464: a ``sudden transition" from one cosmological era to another that is responsible
465: for amplification, but the entire interval of the long-wavelength
466: (non-adiabatic) regime.
467: \par
468: After the end of amplification, the accumulated
469: (and typically large) squeeze parameter $r_n$ stays approximately constant.
470: The mode is again in the adiabatic regime. In course of the evolution, the
471: complex functions
472: ${\stackrel{s}{u}}_n(\eta ) + {\stackrel{s}{v}}_n^{\ast}(\eta )$
473: become practically real, and one has
474: ${\stackrel{s}{h}}_n(\eta ) \approx {\stackrel{s}{h}}_n^{\ast}(\eta ) \approx
475: \frac{1}{a} e^{r_n} \cos \phi_n(\eta)$.
476: Every amplified mode $n$ of the field (\ref{6}) takes the form of
477: a product of a function of time and a (random, operator-valued)
478: function of spatial coordinates; the mode acquires a
479: standing-wave pattern. The periodic dependence $\cos \phi_n(\eta)$
480: will be further discussed below.
481: \par
482: It is clearly seen from the fundamental equations (\ref{10}), (\ref{11}),
483: (\ref{14}) that
484: the final results depend only on $a(\eta)$. Equations do not ask us the
485: names of our favorite cosmological prejudices, they ask us about the
486: pump field $a(\eta)$. Conversely, from the measured relic gravitational waves,
487: we can deduce the behaviour of $a(\eta)$, which is essentially the purpose
488: of detecting the relic gravitons.
489:
490:
491: \section{Cosmological Pump Field}
492:
493:
494: With the chosen initial conditions, the final
495: numerical results for relic gravitational waves depend on the concrete
496: behaviour of the pump
497: field represented by the cosmological scale factor $a(\eta)$. We know
498: a great deal about $a(\eta)$. We know that $a(\eta)$ behaves as
499: $a(\eta) \propto \eta^2$
500: at the present matter-dominated stage. We know that this stage
501: was preceeded by the radiation-dominated stage $a(\eta) \propto \eta$.
502: At these two stages of evolution the functions $a(\eta)$ are simple
503: power-law functions of $\eta$.
504: What we do not know is the function $a(\eta)$ describing the initial stage
505: of expansion of the very early Universe, that is, before
506: the era of primordial nucleosynthesis. It is convenient to
507: parameterize $a(\eta)$ at this initial stage also by power-law functions
508: of $\eta$. First, this is a sufficiently broad class of functions, which,
509: in addition, allows us to find exact solutions to our fundamental
510: equations. Second, it is known \cite{g1} that
511: the pump fields $a(\eta)$ which have power-law dependence in terms of $\eta$,
512: produce gravitational waves with simple power-law spectra in terms of $\nu$.
513: These spectra are easy to analyze and discuss in the context of detection.
514: \par
515: We model cosmological expansion by several successive eras. Concretely,
516: we take $a(\eta)$ at the initial stage of expansion ($i$-stage) as
517: \begin{equation}
518: \label{17}
519: a(\eta) = l_o|\eta|^{1 + \beta},
520: \end{equation}
521: where $\eta$ grows from $- \infty$, and $1 + \beta < 0$. We will show
522: later how the available observational data constrain the
523: parameters $l_o$ and $\beta$. The $i$-stage lasts up to a certain
524: $\eta = \eta_1$, $\eta_1 < 0$. To make our analysis more general,
525: we assume that the $i$-stage was followed by some interval of the
526: $z$-stage ($z$ from Zeldovich). It is known that an
527: interval of evolution governed by the most ``stiff" matter
528: (effective equation of state $p = \epsilon$) advocated by Zeldovich,
529: leads to a relative increase of gravitational wave amplitudes \cite{g1}.
530: It is also known that the requirement of conistency of the graviton production
531: with the observational restrictions does not allow the ``stiff" matter interval
532: to be too much long \cite{g1}, \cite{zn}. However, we want to
533: investigate any interval of
534: cosmological evolution that can be consistently included. In fact,
535: the $z$-stage of expansion that we include is quite general.
536: It can be governed by a ``stiffer than radiation" \cite{gio} matter,
537: as well as by a ``softer than
538: radiation" matter. It can also be simply a part of the radiation-dominated
539: era. Concretely, we take $a(\eta)$ at the interval of time from $\eta_1$
540: to some $\eta_s$ ($z$-stage) in the form
541: \begin{equation}
542: \label{18}
543: a(\eta) = l_o a_z (\eta - \eta_p)^{1 + \beta_s},
544: \end{equation}
545: where $1 + \beta_s > 0$. For the particular choice
546: $\beta_s = 0$, the $z$-stage reduces to an interval
547: of expansion governed by the radiation-dominated matter. Starting
548: from $\eta_s$ and up to $\eta_2$ the Universe was governed by the
549: radiation-dominated matter ($e$-stage). So, at this interval of evolution,
550: we take the scale factor in the form
551: \begin{equation}
552: \label{19}
553: a(\eta) = l_oa_e(\eta - \eta_e).
554: \end{equation}
555: And, finally, from $\eta =\eta_2$ the expansion went over into the
556: matter-dominated era ($m$-stage):
557: \begin{equation}
558: \label{20}
559: a(\eta) = l_oa_m(\eta - \eta_m)^2.
560: \end{equation}
561: A link between the arbitrary constants participating in
562: Eqs. (\ref{17}) - (\ref{20}) is provided
563: by the conditions of continuous joining of the functions $a(\eta)$
564: and $a^{\prime}(\eta)$ at points of transitions $\eta_1$, $\eta_s$, $\eta_2$.
565: \par
566: We denote the present time by $\eta_R$ ($R$ from reception). This time is
567: defined by the observationally known value of the present-day Hubble
568: parameter $H(\eta_R)$ and Hubble radius $l_H=c/{H(\eta_R)}$.
569: For numerical estimates we will be using
570: $l_H \approx 2 \times 10^{28}~{\rm cm}$. It is convenient to choose
571: $\eta_R - \eta_m = 1$, so that $a(\eta_R) = 2l_H$. The ratio
572: \[
573: a(\eta_R)/a(\eta_2) \equiv \zeta_2
574: \]
575: is believed to be around $\zeta_2 = 10^4$. We also denote
576: \[
577: a(\eta_2)/a(\eta_s) \equiv \zeta_s~,~~~ a(\eta_s)/a(\eta_1) \equiv \zeta_1~.
578: \]
579: With these definitions, all the constants participating in
580: Eqs. (\ref{17}) - (\ref{20}) (except
581: parameters $\beta$ and $\beta_s$ which should be chosen from
582: other considerations)
583: are being expressed in terms of $l_H$, $\zeta_2$, $\zeta_s$,
584: and $\zeta_1$. For example,
585: \[
586: |\eta_1|= \frac{|1+\beta|}{2\zeta_2^{\frac{1}{2}}
587: \zeta_s{\zeta_1}^{\frac{1}{1+\beta_s}}} ~.
588: \]
589: The important constant $l_o$ is expressed as
590: \begin{equation}
591: \label{21}
592: l_o = b l_H\zeta_2^{\frac{\beta-1}{2}}\zeta_s^{\beta}{\zeta_1}^{\frac{\beta-\beta_s}{1+\beta_s}} ,
593: \end{equation}
594: where $b \equiv 2^{2+\beta}/{|1+\beta|}^{1+\beta}$. Note that $b=1$ for
595: $\beta = -2$.
596: [This expression for $l_o$ may help to relate formulas written here with
597: the equivalent treatment \cite{g3} which was given in slightly
598: different notations.]
599: The sketch of the entire evolution $a(\eta)$ is given in $Fig. 4$.
600:
601: %Fig. 4
602: \begin{figure}
603: \epsfxsize=0.8\textwidth
604: \centerline{\epsfbox{scalefac.fig4}}
605: \end{figure}
606:
607: \par
608: We work with the spatially-flat models (\ref{4}).
609: At every instant of time, the energy density $\epsilon(\eta)$ of matter
610: driving the evolution is related with the Hubble radius $l(\eta)$ by
611: \begin{equation}
612: \label{22}
613: \kappa \epsilon(\eta) = \frac{3}{l^2(\eta)},
614: \end{equation}
615: where $\kappa = 8\pi G/ c^4$.
616: For the case of power-law scale factors $a(\eta) \propto \eta^{1+\beta}$,
617: the effective pressure $p(\eta)$ of the matter is related with the
618: $\epsilon(\eta)$ by the effective equation of state
619: \begin{equation}
620: \label{23}
621: p = \frac{1-\beta}{3(1+\beta)} \epsilon.
622: \end{equation}
623: For instance, $p=0$ for $\beta =1$, $p=\frac{1}{3} \epsilon$ for $\beta=0$,
624: $p= -\epsilon$ for $\beta = -2$, and so on. Each interval of
625: the evolution (\ref{17})-(\ref{20}) is governed by one of these
626: equations of state.
627: \par
628: In principle, the function $a(\eta)$ could be even more complicated than
629: the one that we consider.
630: It could even include an interval of the early contraction,
631: instead of expansion, leading to the ``bounce" of the scale factor.
632: In case of a decreasing $a(\eta)$ the gravitational-wave equation can
633: still be analyzed and the amplification
634: is still effective \cite{g1}. However, the Einstein equations for
635: spatially-flat models do not permit a
636: regular ``bounce" of $a(\eta)$ (unless $\epsilon$ vanishes at the moment
637: of ``bounce"). Possibly, a ``bounce" solution can be realized in alternative
638: theories, such, for example, as string-motivated cosmologies
639: \cite{vg}. For a
640: recent discussion of spectral slopes of gravitational waves produced
641: in ``bounce" cosmologies, see~\cite{cr}.
642:
643:
644:
645:
646: \section{Solving Gravitational Wave Equations}
647:
648:
649: The evolution of the scale factor $a(\eta)$ given by
650: Eqs. (\ref{17}) - (\ref{20}) and
651: sketched in $Fig. 4$ allows us to calculate the function $a'/a$. This function
652: is sketched in $Fig. 5$. In all the theoretical generality, the left-hand-side
653: of the barrier in $Fig. 5$ could also consist of several pieces, but we
654: do not consider this possibility here. The graph also shows the
655: important wave numbers
656: $n_H$, $n_2$, $n_s$, $n_1$. The $n_H$ marks the wave whose today's wavelength
657: $\lambda(\eta_R)= 2 \pi a(\eta_R)/n_H$ is equal to the today's Hubble
658: radius $l_H$. With our parametrization $a(\eta_R) = 2 l_H$, this wavenumber
659: is $n_H = 4 \pi$. The $n_2$ marks the wave whose wavelength $\lambda(\eta_2)=
660: 2 \pi a(\eta_2)/n_2$ at $\eta= \eta_2$ is equal to the Hubble
661: radius $l(\eta_2)$ at $\eta = \eta_2$. Since
662: $\lambda(\eta_R)/\lambda(\eta_2) = (n_2/n_H)[a(\eta_R)/a(\eta_2)]$ and
663: $l(\eta_R)/l(\eta_2) = [a(\eta_R)/a(\eta_2)][a(\eta_R)/a(\eta_2)]^{1/2}$,
664: this gives us $n_2/n_H = [a(\eta_R)/a(\eta_2)]^{1/2} = \zeta_2^{1/2}$.
665: Working out in a similar fashion other ratios, we find
666: \begin{equation}
667: \label{24}
668: \frac{n_2}{n_H} = \zeta_2^{\frac{1}{2}},~~~ \frac{n_s}{n_2} = \zeta_s,
669: ~~~\frac{n_1}{n_s} =\zeta_1~^{\frac{1}{1+\beta_s}}.
670: \end{equation}
671:
672:
673:
674: \par
675: Solutions to the gravitational wave equations exist for any $a(\eta)$.
676: At intervals of power-law dependence $a(\eta)$, solutions to Eq. (\ref{11})
677: have simple form of the Bessel functions. We could have found piece-wise
678: exact solutions to Eq. (\ref{11}) and join them in the transition points.
679: However, we will use a much simpler treatment which is sufficient for our
680: purposes. We know that the
681: squeeze parameter $r_n$ stays constant in the short-wavelength regimes and
682: grows according to Eq. (\ref{15}) in the long-wavelength regime. All modes
683: start in the vacuum state, that is, $r_n = 0$ initially. After the end of
684: amplification, the accumulated value (\ref{16}) stays constant up to today.
685: To find today's value of $e^{r_n}$ we need to calculate the
686: ratio $a_{**}(n)/a_*(n)$. For
687: every given $n$, the quantity $a_*$ is determined by the condition
688: $\lambda(\eta_*) = l(\eta_*)$, wheras $a_{**}$ is determined by the
689: condition $\lambda(\eta_{**}) = l(\eta_{**})$.
690:
691: %Fig. 5
692: \begin{figure}
693: \epsfxsize=0.8\textwidth
694: \centerline{\epsfbox{aprime-a.fig5}}
695: \end{figure}
696:
697: \par
698: Let us start from the mode $n = n_1$. For this wave number we have
699: $a_* = a_{**} = a (\eta_1)$, and therefore $r_{n_1} = 0$. The higher frequency
700: modes $n > n_1$ (above the barrier in $Fig. 5$) have never been in the
701: amplifying regime, so we can write
702: \begin{equation}
703: \label{25}
704: e^{r_n} = 1,~~ n\ge n_1.
705: \end{equation}
706: Let us now consider the modes $n$ in the interval $n_1 \ge n \ge n_s$.
707: For a given $n$ we need to know $a_* (n)$ and $a_{**}(n)$. Using Eq. (\ref{17})
708: one finds $a_*(n)/a_*(n_1) = (n_1/n)^{1 +\beta}$, and using Eq. (\ref{18}) one
709: finds $a_{**}(n)/a_{**}(n_s) = (n_s/n)^{1 +\beta_s}$. Therefore, one finds
710: \[
711: \frac{a_{**}(n)}{a_*(n)} = \frac{a_{**}(n_s)}{a_*(n_1)}
712: \left(\frac{n_s}{n}\right)^{1+\beta_s}\left(\frac{n}{n_1}\right)^{1+\beta}.
713: \]
714: Since $a_{**}(n_s) = a(\eta_s)$, $a_*(n_1) = a(\eta_1)$, and
715: $a(\eta_s)/a(\eta_1) = \zeta_1 =(n_1/n_s)^{1+\beta_s}$, we arrive at
716: \[
717: \frac{a_{**}(n)}{a_*(n)} = \left(\frac{n}{n_1}\right)^{\beta - \beta_s}.
718: \]
719: Repeating this analysis for other intervals of the decreasing $n$,
720: we come to the conclusion that
721: \begin{eqnarray}
722: \label{26}
723: e^{r_n}& =& \left(\frac{n}{n_1}\right)^{\beta - \beta_s},~~
724: n_1 \ge n \ge n_s, \nonumber \\
725: e^{r_n}& =& \left(\frac{n}{n_s}\right)^{\beta}
726: \left(\frac{n_s}{n_1}\right)^{\beta - \beta_s} ,~~
727: n_s \ge n \ge n_2, \nonumber \\
728: e^{r_n}& =& \left(\frac{n}{n_2}\right)^{\beta - 1}
729: \left(\frac{n_2}{n_1}\right)^{\beta}\left(\frac{n_s}{n_1}\right)^{-\beta_s},
730: n_2 \ge n \ge n_H.
731: \end{eqnarray}
732: The mnemonic rule of constructing $e^{r_n}$ at successive intervals
733: of decreasing $n$ is simple.
734: If the interval begins at $n_x$, one takes $(n/n_x)^{\beta_{*} - \beta_{**}}$
735: and multiples with $e^{r_{n_x}}$, that is, with the previous interval's
736: value $e^{r_n}$ calculated at the end of that interval $n_x$.
737: For the function $a'/a$ that we are working with, the $\beta_{*}$ is always
738: $\beta$, whereas the $\beta_{**}$ takes the values $\beta_s$, $0$, $1$ at
739: the successive intervals.
740: \par
741: The modes with $n < n_H$ are still in the long-wavelength regime. For these
742: modes, we should take $a(\eta_R)$ instead of $a_{**}(n)$. Combining with
743: $a_*(n)$, we find
744: \begin{equation}
745: \label{27}
746: e^{r_n} = \left(\frac{n}{n_H}\right)^{\beta +1}
747: \left(\frac{n_H}{n_2}\right)^{\beta - 1}
748: \left(\frac{n_2}{n_1}\right)^{\beta}\left(\frac{n_s}{n_1}\right)^{-\beta_s},
749: ~~ n \le n_H.
750: \end{equation}
751: Formulas (\ref{25}) - (\ref{27}) give approximate values of $r_n$
752: for all $n$. The
753: factor $e^{r_n}$ is $e^{r_n} \ge 1$ for $n \le n_1$, and
754: $e^{r_n} \gg 1$ for $n \ll n_1$. This factor determines the mean square
755: amplitude of the gravitational waves.
756: \par
757: The mean value of the field $h_{ij}$ is zero at every moment of time $\eta$
758: and in every spatial point ${\bf x}$:
759: $\langle 0|h_{ij}(\eta, {\bf x})|0\rangle = 0$.
760: The variance
761: \[
762: \langle 0|h_{ij}(\eta, {\bf x})h^{ij}(\eta, {\bf x})|0\rangle ~\equiv~
763: \langle h^2 \rangle
764: \]
765: is not zero, and it determines the mean square amplitude of the generated
766: field - the quantity of interest for the experiment. Taking the product
767: of two expressions (\ref{6}) one can show that
768: \begin{equation}
769: \label{28}
770: \langle h^2 \rangle = \frac{C^2}{2\pi^2} \int_0^\infty n \sum_{s=1}^2
771: \Big| {\stackrel{s}{h}}_n(\eta )\Big|^2 ~{\rm d}n \equiv
772: \int_0^\infty h^2(n, \eta) \frac{{\rm d}n}{n}~.
773: \end{equation}
774: Using the representation (\ref{7}), (\ref{8}) in Eq. (\ref{28}) one can
775: also write
776: \begin{equation}
777: \label{29}
778: \langle h^2 \rangle =
779: \frac{C^2}{{\pi^2}a^2 } \int_0^\infty n {\rm d}n (\cosh2{r_n} +
780: \cos2{\phi_n}\sinh2{r_n}).
781: \end{equation}
782: We can now consider the present era and use the fact that $e^{r_n}$ are
783: large numbers for all $n$ in the interval of our interest
784: $n_1 \ge n \ge n_H$. Then, we can derive
785: \begin{equation}
786: \label{30}
787: h(n, \eta)\approx \frac{C}{\pi}\frac{1}{a(\eta_R)}n e^{r_n}\cos\phi_n(\eta)=
788: 8 \sqrt{\pi}\left(\frac{l_{Pl}}{l_H}\right)\left(\frac{n}{n_H}\right)
789: e^{r_n}\cos\phi_n(\eta) ~.
790: \end{equation}
791: The quantity $h(n, \eta)$ is the
792: dimensionless spectral amplitude of the field whose numerical value
793: is determined by the calculated squeeze parameter $r_n$.
794: The oscillatory factor $\cos\phi_n(\eta)$ reflects the squeezing (standing
795: wave pattern) acquired by modes with $n_1 > n > n_H$. For modes with
796: $n < n_H$ this factor is approximately $1$. For high-frequency modes
797: $n \gg n_H$ one has
798: $\phi_n(\eta) \approx n(\eta - \eta_n) \gg 1$, so that $h(n, \eta)$ makes
799: many oscillations while the scale factor $a(\eta)$ is practically fixed
800: at $a(\eta_R)$.
801: \par
802: The integral (\ref{29}) extends formally from $0$ to $\infty$.
803: Since $r_n \approx 0$
804: for $n \ge n_1$, the integral diverges at the upper limit. This is a typical
805: ultra-violet divergence. It should be discarded (renormalized to zero)
806: because it comes from
807: the modes which have always been in their vacuum state. At the lower limit,
808: the integral diverges, if $\beta \le -2$. This is an infra-red divergence
809: which comes from the assumption that the amplification process has started
810: from infinitely remote time in the past. One can deal with this
811: divergence either by introducing a lower frequency cut-off (equivalent to
812: the finite duration of the amplification) or by considering only the
813: parameters $\beta > -2$, in which case the integral is convergent at the lower
814: limit.
815: It appears that the available observational data (see below) favour this
816: second option. The particular case $\beta = -2$ corresponds to
817: the de Sitter evolution $a(\eta) \propto |\eta|^{-1}$. In this case,
818: the $h(n)$ found in Eqs. (\ref{30}), (\ref{27}) does not depend on $n$.
819: This is known as the Harrison-Zeldvich, or scale-invariant, spectrum.
820: \par
821: An alternative derivation of the spectral amplitude $h(n)$ uses the
822: approximate solutions (\ref{12}), (\ref{13}) to the wave equation (\ref{11}).
823: This method gives exactly the same, as in Eqs. (\ref{30}),
824: (\ref{25}) - (\ref{27}) numerical values of $h(n)$, but does not
825: reproduce the oscillatory factor $\cos\phi_n(\eta)$.
826: \par
827: One begins with the initial spectral amplitude $h_i(n)$ defined
828: by quantum normalization:
829: $h_i(n) = 8 \sqrt{\pi} (l_{Pl}/ \lambda_i)$. This is the amplitude of the mode
830: $n$ at the moment $\eta_*$ of entering the long wavelength regime, i.e.
831: when the mode's wavelength $\lambda_i$ is equal to the Hubble
832: radius $l(\eta_*)$. For $\lambda_i$ one derives
833: \begin{equation}
834: \label{31}
835: \lambda_i = \frac{1}{b} l_o \left(\frac{n_H}{n}\right)^{2+\beta}.
836: \end{equation}
837: Thus, we have
838: \begin{equation}
839: \label{32}
840: h_i(n) = A \left(\frac{n}{n_H}\right)^{2+\beta},
841: \end{equation}
842: where $A$ denotes the constant
843: \begin{equation}
844: \label{33}
845: A = b 8\sqrt{\pi} \frac{l_Pl}{l_o}~.
846: \end{equation}
847: The numbers $h_i(n)$ are defined at the beginning of the long-wavelength
848: regime. In other words, they are given along the left-hand-side slope of the
849: barrier in $Fig. 5$.
850: We want to know the final numbers (spectral amplitudes) $h(n)$
851: which describe the field today, at $\eta_R$.
852: \par
853: According to the dominant solution $h_n(\eta) =const$ of the long-wavelength
854: regime (see Eq. (\ref{d})), the initial amplitude $h_i(n)$ stays practically
855: constant up to the end of the long-wavelength regime at $\eta_{**}$,
856: that is, up to the right-hand-side slope of the barrier.
857: [The second term in Eq. (\ref{13}) could be
858: important only at the $z$-stage and only for parameters
859: $\beta_s \le -(1/2)$, which
860: correspond to the effective equations of state $p \ge \epsilon$.
861: In order to keep the analysis simple, we do not consider those cases.]
862: After the completion of the long-wavelength regime, the amplitudes decrease
863: adiabatically in proportion to $1/a(\eta)$, up to the present time.
864: Thus, we have
865: \begin{equation}
866: \label{34}
867: h(n) = A \left(\frac{n}{n_H}\right)^{2+\beta} \frac{a_{**}(n)}{a(\eta_R)}.
868: \end{equation}
869: \par
870: Let us start from the lower end of the spectrum, $n \le n_H$, and go
871: upward in $n$. The modes $n \le n_H$ have not started yet the adiabatic
872: decrease of the amplitude, so we have
873: \begin{equation}
874: \label{35}
875: h(n) = A \left(\frac{n}{n_H}\right)^{2+\beta},~~ n \le n_H.
876: \end{equation}
877: Now consider the interval $n_2 \ge n \ge n_H$. At this interval,
878: the $a_{**}(n)/a(\eta_R)$ scales as $(n_H/n)^2$, so we have
879: \begin{equation}
880: \label{36}
881: h(n) = A \left(\frac{n}{n_H}\right)^{\beta},~~ n_2 \ge n \ge n_H.
882: \end{equation}
883: At the interval $n_s \ge n \ge n_2$ the ratio
884: $a_{**}(n)/a(\eta_R) = [a_{**}(n)/a(\eta_2)][a(\eta_2)/a(\eta_R)]$
885: scales as $(n_2/n)(n_H/n_2)^2$, so we have
886: \begin{equation}
887: \label{37}
888: h(n) =A\left(\frac{n}{n_H}\right)^{1+\beta}\frac{n_H}{n_2},~~ n_s\ge n\ge n_2.
889: \end{equation}
890: Repeating the same analysis for the interval $n_1 \ge n \ge n_s$ we find
891: \begin{equation}
892: \label{38}
893: h(n) =A\left(\frac{n}{n_H}\right)^{1+\beta -\beta_s}
894: \left(\frac{n_s}{n_H}\right)^{\beta_s}\frac{n_H}{n_2},~~ n_1\ge n\ge n_s.
895: \end{equation}
896: It is seen from
897: Eq. (\ref{38}) that an interval of the $z$-stage with $\beta_s <0$
898: (the already imposed
899: restrictions require also $(-1/2) < \beta_s$) bends
900: the spectrum $h(n)$ upwards, as compared with Eq. (\ref{37}), for larger $n$.
901: If one recalls the relationship (\ref{21}) between $l_o$ and $l_H$
902: and uses (\ref{26}), (\ref{27}) in Eq. (\ref{30}) one
903: arrives exactly at Eqs. (\ref{35})-(\ref{38}) up to
904: the oscillating factor $\cos\phi_n(\eta)$.
905: \par
906: Different parts of the barrier in $Fig. 5$ are responsible for amplitudes and
907: spectral slopes at different intervals of $n$. The sketch of the generated
908: spectrum $h(n)$ in conjunction with the form of the barrier
909: is shown in $Fig. 6$.
910:
911:
912: %Fig. 6
913: \begin{figure}
914: \epsfxsize=0.8\textwidth
915: \centerline{\epsfbox{charam.fig6}}
916: \end{figure}
917:
918:
919: The present day frequency of the oscillating modes, measured in $Hz$,
920: is defined as $\nu = cn/2\pi a(\eta_R)$. The lowest frequency (Hubble
921: frequency) is $\nu_H = c/l_H$. For numerical estimates we will be using
922: $\nu_H \approx 10^{-18} Hz$. The ratios of $n$ are equal to the ratios of
923: $\nu$, so that, for example, $n/n_H = \nu/\nu_H$. For high-frequency
924: modes we will now often use the ratios of $\nu$ instead of ratios of $n$.
925: \par
926: In addition to the spectral amplitudes $h(n)$ the generated field can be
927: also characterized by the spectral energy density parameter $\Omega_{g}(n)$.
928: The energy density $\epsilon_g$ of the gravitational wave field is
929: \[
930: \kappa \epsilon_g = \frac{1}{4} h^{ij}_{~,0} h_{ij,0} =
931: \frac{1}{4a^2} {h^{ij}}^{\prime} {h_{ij}}^{\prime}.
932: \]
933: The mean value $\langle 0|\epsilon_g (\eta, {\bf x})|0\rangle$ is given by
934: \begin{equation}
935: \label{39}
936: \kappa \langle \epsilon_g \rangle = \frac{1}{4a^2}
937: \frac{C^2}{2\pi^2} \int_0^\infty n \sum_{s=1}^2
938: \Big| {{\stackrel{s}{h}}}^{\prime}_n(\eta )\Big|^2 ~{\rm d}n.
939: \end{equation}
940: For high-frequency modes, it is only the factor $e^{\pm i n\eta}$ that needs
941: to be differentiated by $\eta$. After avaraging out the oscillating factors,
942: one gets
943: $\Big| {{\stackrel{s}{h}}}^{\prime}_n\Big|^2 =
944: n^2 \Big| {\stackrel{s}{h}}_n\Big|^2$, so that
945: \begin{equation}
946: \label{40}
947: \kappa \langle \epsilon_g \rangle = \frac{1}{4a^2}
948: \int_0^\infty n^2 h^2(n) \frac{{\rm d}n}{n}.
949: \end{equation}
950: In fact, the high-frequency approximation, that has been used, permits
951: integration over lower $n$ only up to $n_H$. And the upper limit, as was
952: discussed above, is in practice $n_1$, not infinity.
953: The parameter $\Omega_g$ is defined as $\Omega_g =
954: \langle \epsilon_g \rangle /\epsilon$, where $\epsilon$ is given by
955: Eq. (\ref{22}) (critical density). So, we derive
956: \[
957: \Omega_g = \int_{n_H}^{n_1} \Omega_g(n) \frac{{\rm d}n}{n} =
958: \int_{\nu_H}^{\nu_1} \Omega_g(\nu) \frac{{\rm d}\nu}{\nu}
959: \]
960: and
961: \begin{equation}
962: \label{41}
963: \Omega_g(\nu) = \frac{\pi^2}{3}h^2(\nu)\left(\frac{\nu}{\nu_H}\right)^2.
964: \end{equation}
965: \par
966: The dimensionless quantity $\Omega_g(\nu)$ is useful because it allows us
967: to quickly evaluate the cosmological importance of the generated field in
968: a given frequency interval. However, the primary and more universal concept
969: is $h(\nu)$, not $\Omega_g(\nu)$. It is the field, not its energy density,
970: that is directly measured by the gravity-wave detector. One should also note
971: that some authors use quite a misleading definition $\Omega_g (f) =
972: (1/\rho_c)(d \rho_{gw}/d \ln f)$ which suggests differentiation of the
973: gravity-wave energy density by frequency. This would be incorrect and could
974: cause disagreements in numerical values of $\Omega_g$. Whenever we use
975: $\Omega_g(\nu)$, we mean relationship (\ref{41}); and for order
976: of magnitude estimates one can use \cite{g1}:
977: \begin{equation}
978: \label{42}
979: \Omega_g(\nu) \approx h^2(\nu) \left(\frac{\nu}{\nu_H}\right)^2 .
980: \end{equation}
981:
982:
983:
984: \section{Theoretical and Observational Constraints}
985:
986:
987:
988: The entire theoretical approach is based on the assumption that a weak
989: quantized gravity-wave field interacts with a classical pump field. We
990: should follow the validity of this approximation throughout the analysis.
991: The pump field can be treated as a classical gravitational field
992: as long as the driving energy
993: density $\epsilon$ is smaller than the Planck energy density, or, in
994: other words, as long as the Hubble radius $l(\eta)$ is greater than the
995: Planck length $l_{Pl}$. This is a restriction on the pump field, but
996: it can be used as a restriction on the
997: wavelength $\lambda_i$ of the gravity-wave mode $n$ at the
998: time of entry the long-wavelength regime. If $l(\eta_*) > l_{Pl}$, then
999: $\lambda_i > l_{Pl}$. The $\lambda_i$ is given by Eq. (\ref{31}).
1000: So, we need to ensure that
1001: \[
1002: b \frac{l_{Pl}}{l_o} \left(\frac{\nu}{\nu_H}\right)^{2+\beta} < 1.
1003: \]
1004: At the lowest-frequency end $\nu = \nu_H$ this inequality
1005: gives $b(l_{Pl}/l_o) < 1$. In fact, the observational constraints
1006: (see below) give a stronger restiction:
1007: \begin{equation}
1008: \label{43}
1009: b \frac{l_{Pl}}{l_o} \approx 10^{-6} ,
1010: \end{equation}
1011: which we accept. Then, at the
1012: highest-frequency end $\nu = \nu_1$ we need to satisfy
1013: \begin{equation}
1014: \label{44}
1015: \left(\frac{\nu_1}{\nu_H}\right)^{2+\beta} < 10^6.
1016: \end{equation}
1017: \par
1018: Let us now turn to the generated spectral amplitudes $h(\nu)$. According
1019: to Eq. (\ref{35}) we have $h(\nu_H) \approx b 8 \sqrt{\pi} (l_{Pl}/l_o)$. The
1020: measured microwave beckgound anisotropies, which we discuss below,
1021: require this number
1022: to be at the level of $10^{-5}$, which gives the already mentioned
1023: Eq. (\ref{43}). The quantity $h(\nu_1)$ at the highest
1024: frequency $\nu_1$ is given by Eq. (\ref{38}):
1025: \[
1026: h(\nu_1) = b 8\sqrt{\pi} \frac{l_{Pl}}{l_o}
1027: \left(\frac{\nu_1}{\nu_H}\right)^{1+\beta -\beta_s}
1028: \left(\frac{\nu_s}{\nu_H}\right)^{\beta_s}\frac{\nu_H}{\nu_2}.
1029: \]
1030: Using Eq. (\ref{21}) this expression for $h(\nu_1)$ can be rewritten as
1031: \begin{equation}
1032: \label{45}
1033: h(\nu_1) =8\sqrt{\pi} \frac{l_{Pl}}{l_H}\frac{\nu_1}{\nu_H}=
1034: 8\sqrt{\pi} \frac{l_{Pl}}{\lambda_1},
1035: \end{equation}
1036: where $\lambda_1 = c/\nu_1$. This last expression for $h(\nu_1)$ is not
1037: surprising: the modes with $\nu \ge \nu_1$ are still in the vacuum state,
1038: so the numerical value of $h(\nu_1)$ is determined by quantum normalization.
1039: \par
1040: All the amplified modes have started with small initial amplitudes
1041: $h_i$, at the level of
1042: zero-point quantum fluctuations. These amplitudes are also small today, since
1043: the $h_i$ could only stay constant or decrease. However, even these relatively
1044: small amplitudes should obey observational constraints. We do not want the
1045: $\Omega_g$ in the high-frequency modes, which might affect the rate of
1046: the primordial nucleosynthesis, to exceed the level of $10^{-5}$. This
1047: means that $\Omega_g(\nu_1)$ cannot exceed the level of $10^{-6}$ or so.
1048: The use of Eq. (\ref{41}) in combination with
1049: $\Omega_g(\nu_1) \approx 10^{-6}$ and
1050: $h(\nu_1)$ from Eq. (\ref{45}), gives us the highest allowed frequency
1051: $\nu_1 \approx 3 \times 10^{10} Hz$. We will use this value of $\nu_1$
1052: in our numerical estimates.
1053: Returning with this value of $\nu_1$ to Eq. (\ref{44}) we find that parameter
1054: $\beta$ can only be $\beta \le - 1.8$. We will be treating $\beta = -1.8$
1055: as the upper limit for the allowed values of $\beta$.
1056: \par
1057: We can now check whether the accepted parameters leave room for the
1058: postulated $z$-stage with $\beta_s < 0$. Using Eq. (\ref{21}) we can rewrite
1059: Eq. (\ref{43}) in the form
1060: \begin{equation}
1061: \label{46}
1062: 10^{-6} \frac{l_H}{l_{Pl}} =
1063: \left(\frac{\nu_1}{\nu_H}\right)^{- \beta}
1064: \left(\frac{\nu_1}{\nu_s}\right)^{\beta_s}\frac{\nu_2}{\nu_H}.
1065: \end{equation}
1066: We know that $\nu_2/ \nu_H = 10^2$ and $\nu_1/\nu_s$ is not smaller
1067: than $1$. Substituting all the numbers in Eq. (\ref{46}) one can find that
1068: this equation cannot be satisfied for the largest possible $\beta = -1.8$.
1069: In the case $\beta = -1.9$,
1070: Eq. (\ref{46}) is only marginally satisfied, in the sense that a significant
1071: deviation from $\beta_s =0$ toward negative $\beta_s$ can only last for
1072: a relatively short time. For
1073: instance, one can accomodate $\beta_s = -0.4$ and $\nu_s = 10^8 Hz$.
1074: On the other hand, if one takes
1075: $\beta = -2$, a somewhat longer interval of the $z$-stage with $\beta_s < 0$
1076: can be
1077: included. For instance, Eq. (\ref{46}) is satisfied if one accepts $\nu_s =
1078: 10^{-4} Hz$ and $\beta_s = - 0.3$. This allows us to slightly increase $h(\nu)$
1079: in the interval $\nu_s < \nu < \nu_1$, as compared with the values of
1080: $h(\nu)$ reached in the more traditional
1081: case $\beta = -2$, $\beta_s = 0$. In what follows, we will consider
1082: consequences of this assumption for the prospects of detection of the
1083: produced gravitational wave signal.
1084: Finally, let us see what the available information on the microwave background
1085: anisotropies \cite{s}, \cite{b} allows us to conclude about the
1086: parameters $\beta$ and $l_o$.
1087: \par
1088: Usually, cosmologists operate with the spectral index ${\rm n}$ (not to be
1089: confused with the wave number $n$) of primordial cosmological perturbations.
1090: Taking into account the way in which the spectral index ${\rm n}$ is defined,
1091: one can relate ${\rm n}$ with the spectral index $\beta + 2$ that shows up
1092: in Eq. (\ref{35}). The relationship between them is ${\rm n} = 2\beta+5$.
1093: This relationship is valid independently of the nature of cosmological
1094: perturbations. In particular, it is valid for density perturbations,
1095: in which case the $h(n)$ of Eq. (\ref{35}) is the dimensionless
1096: spectral amplitude
1097: of metric perturbations associated with density perturbations. If primordial
1098: gravitational waves and density perturbations were generated by the
1099: mechanism that
1100: we discuss here (the assumption that is likely to be true) than the parameter
1101: $\beta$ that participates in the spectral index is the same one that
1102: participates in the scale factor of Eq. (\ref{17}).
1103: Primordial gravitational waves and primordial density perturbations
1104: with the same spectral index produce approximately the same
1105: lower-order multipole distributions of large-scale anisotropies.
1106: \par
1107: The evaluation of the spectral
1108: index ${\rm n}$ of primordial perturbations have resulted
1109: in ${\rm n} = 1.2 \pm 0.3$ \cite{b} or even in a somewhat higher
1110: value. A recent analysis \cite{melch} of all
1111: available data favors ${\rm n} = 1.2$ and the quadrupole contribution of
1112: gravitational waves twice as large as that of density perturbations.
1113: One can interpret these evaluations
1114: as indication that the true value of ${\rm n}$ lies somewhere near
1115: ${\rm n} = 1.2$ (hopefully, the planned new observational missions will
1116: determine this index more accurately). This gives us the parameter $\beta$
1117: somewhere near $\beta = -1.9$. We will be using $\beta = -1.9$ in our estimates
1118: below, as the observationally preferred value. The parameter $\beta$ can
1119: be somewhat larger than $\beta = -1.9$. However, as we already
1120: discussed, the value $\beta = -1.8$ (${\rm n} = 1.4$)
1121: is the largest one for which the entire approch is well posed.
1122: The Harrison-Zeldovich spectral index ${\rm n} = 1$ corresponds to $\beta = -2$.
1123: \par
1124: The observed quadrupole anisotropy of the microwave background radiation is
1125: at the level $\delta T/ T \approx 10^{-5}$.
1126: The quadrupole anisotropy that would be produced by the spectrum
1127: (\ref{35}) - (\ref{38}) is
1128: mainly accounted for by the wave numbers near $n_H$. Thus,
1129: the numerical value of the quadrupole anisotropy produced by relic
1130: gravitational waves is approximately equal to $A$. According to general
1131: physical considerations and detailed calculations \cite{gden}, the
1132: metric amplitudes of long-wavelength gravitational waves and density
1133: perturbations generated by the discussed amplification mechanism are of the
1134: same order of magnitude. Therefore, they contribute roughly equally to the
1135: anisotropy at lower multipoles. This gives us the estimate $A\approx 10^{-5}$,
1136: that we have already used in Eq. (\ref{43}). It is not yet proven
1137: observationally that a significant part of the observed anosotropies
1138: at lower multipoles is indeed provided by
1139: relic gravitational waves, but we can at least assume this with some degree of
1140: confidence. It is likely that the future measurements of the microwave
1141: background radiation will help us to verify this theoretical conclusion.
1142: \par
1143: Combining all the evaluated parameters together, we show in $Fig. 7$ the
1144: expected spectrum of $h(\nu)$ for the case $\beta = -1.9$. A small allowed
1145: interval of the $z$-stage is also included. The intervals of the spectrum
1146: accessible to space-based and ground-based interferometers are indicated
1147: by vertical lines.
1148:
1149:
1150: %Fig. 7
1151: \begin{figure}
1152: \epsfxsize=0.8\textwidth
1153: \centerline{\epsfbox{predspec.fig7}}
1154: \end{figure}
1155:
1156:
1157: \par
1158: It is necessary to note \cite{gden}, \cite{g3}
1159: that the confirmation of any ${\rm n} > 1$ ($\beta > -2$) would
1160: mean that the
1161: very early Universe was not driven by a scalar field - the cornerstone of
1162: inflationary considerations. This is because the ${\rm n} > 1$ ($\beta > -2$)
1163: requires the effective equation of state at the initial stage of expansion
1164: to be $\epsilon + p < 0$ (see Eq. (\ref{23})), but this cannot be accomodated
1165: by any scalar field
1166: with whichever scalar field potential. The available data do not
1167: prove yet that ${\rm n} > 1$, but this possibility seems likely.
1168: \par
1169: It is also
1170: necessary to say that a certain damage to gravitational wave research
1171: was inflicted by the so called ``standard inflationary result". The
1172: ``standard inflationary
1173: result" predicts infinitely large amplitudes of density perturbations in
1174: the interval of spectrum with the Harrison-Zeldovich
1175: slope ${\rm n} =1$ ($\beta = -2$):
1176: $\delta\rho/\rho \propto 1/\sqrt{1 -{\rm n}}$. The metric (gravitational
1177: field) amplitudes of density perturbations are
1178: also predicted to be infinitely large, in the same proportion.
1179: Through the so-called
1180: ``consistency relation" this divergence leads to the vanishingly small
1181: amplitudes of relic gravitational waves. Thus, the ``standard" inflationary
1182: theory predicts zero for relic gravitational waves; the spectrum similar in
1183: shape to the one shown in $Fig. 7$ would have been shifted down by many
1184: orders of magnitude.
1185: This prediction is hanging on the ``standard inflationary result", but
1186: the ``result" itself is in a severe conflict not only with theory but
1187: with observations too: when the
1188: observers marginalize their data to ${\rm n} = 1$ (enforce this value
1189: of ${\rm n}$ in data analysis) they find finite and small density
1190: perturbations instead of infinitely large perturbations predicted by
1191: inflationary theorists.
1192: [For analytical expressions of the ``standard inflationary result" see any
1193: inflationary article, including recent reviews. For graphical illustration
1194: of the predicted divergent density perturbations and quadrupole anisotropies
1195: see, for example, \cite{ms}. For critical analysis and disagreement with
1196: the ``standard inflationary result" see \cite{gden}.] General relativity and
1197: quantum field theory do not produce the ``standard inflationary result",
1198: so we shall better return to what they say.
1199:
1200:
1201: \section{Detectability of Relic Gravitational Waves}
1202:
1203:
1204: We switch now from cosmology to prospects of detecting the predicted relic
1205: gravitational waves. The ground-based \cite{Abr}-\cite{HoughD}
1206: and space-based \cite{lisa}, \cite{Hel} laser interferometers
1207: (see also \cite{gw0}-\cite{gw2})
1208: will be in the focus of our attention. We use laboratory frequencies
1209: $\nu$ and intervals of laboratory time $t$
1210: $(c{\rm d}t = a(\eta_R){\rm d}\eta)$. Formulas (\ref{37}) and (\ref{38}),
1211: with $A = 10^{-5}$, $\nu_2/\nu_H = 10^2$, and the oscillating factor
1212: restored, can be written as
1213: \begin{equation}
1214: \label{47}
1215: h(\nu,t) \approx
1216: 10^{-7}\cos[2\pi\nu(t - t_{\nu})]\left(\frac{\nu}{\nu_H}\right)^{\beta+1},
1217: ~~\nu_2 \le \nu \le \nu_s
1218: \end{equation}
1219: and
1220: \begin{equation}
1221: \label{48}
1222: h(\nu,t) \approx
1223: 10^{-7}\cos[2\pi\nu(t - t_{\nu})]
1224: \left(\frac{\nu}{\nu_H}\right)^{1+\beta -\beta_s}
1225: \left(\frac{\nu_s}{\nu_H}\right)^{\beta_s}.~~\nu_s \le \nu \le \nu_1
1226: \end{equation}
1227: where the deterministic (not random) constant $t_{\nu}$
1228: does not vary significantly from one frequency to another at
1229: the intervals $\Delta\nu \approx \nu$.
1230: The explicit time dependence of the spectral variance $h^{2}(\nu,t)$
1231: of the field, or, in other words, the explicit time dependence of
1232: the (zero-lag) temporal correlation function of the field at every
1233: given frequency, demonstrates
1234: that we are dealing with a non-stationary process (a consequence
1235: of squeezing and severe reduction of the phase uncertainty).
1236: We will first ignore the oscillating factor and will compare the predicted
1237: amplitudes with the sensitivity curves of advanced detectors. The potential
1238: reserve of improving the signal to noise ratio by expoloiting the squeezing
1239: will be discussed later.
1240: \par
1241: Let us start from the Laser Interferometer Space Antenna (LISA) \cite{lisa}.
1242: The instrument will be most sensitive in the interval, roughly, from
1243: $10^{-3} Hz$ to $10^{-1} Hz$, and will be reasonably sensitive
1244: in a broader range, up to frequencies
1245: $10^{-4} Hz$ and $1 Hz$. The sensitivity graph of LISA to a stochastic
1246: background is usually plotted under the assumption of a 1-year observation
1247: time, that is, the root-mean-square (r.m.s.) instrumental noise is
1248: being evaluated in frequency
1249: bins $\Delta \nu = 3\times 10^{-8} Hz$ around each frequency $\nu$.
1250: We need to rescale our predicted amplitude $h(\nu)$ to these bins.
1251: \par
1252: The mean square amplitude of the gravitational wave field is given by
1253: the integral (\ref{28}). Thus, the r.m.s. amplitude in the band $\Delta \nu$
1254: centered at a given frequency $\nu$ is given by the expression
1255: \begin{equation}
1256: \label{49}
1257: h(\nu, \Delta\nu) = h(\nu) \sqrt{\frac{\Delta\nu}{\nu}}.
1258: \end{equation}
1259: We use Eqs. (\ref{47}), (\ref{48}) and calculate expression (\ref{49})
1260: assuming $\Delta \nu = 3\times 10^{-8} Hz$. The results are plotted
1261: in $Fig. 8$.
1262: Formula (\ref{47}) has been used
1263: throughout the covered frequency interval for the realistic
1264: case $\beta = -1.9$ and for the extreme case $\beta = -1.8$.
1265: The line marked $z$-model describes the signal produced in the
1266: composite model with $\beta = -2$ up to $\nu_s = 10^{-4} Hz$
1267: (formula (\ref{47})) and then followed by formula (\ref{48})
1268: with $\beta_s = -0.3$. This
1269: model gives the signal a factor of $3$ higher at $\nu= 10^{-3} Hz$,
1270: than the model $\beta = -2$ extrapolated down to this frequency.
1271: %Fig.8
1272: \begin{figure}
1273: \epsfxsize=0.8\textwidth
1274: \centerline{\epsfbox{lisa.fig8}}
1275: \end{figure}
1276: \par
1277: There is no doubt that the signal $\beta = -1.8$ would be easily
1278: detectable even with a single instrument. The signal $\beta = -1.9$
1279: is marginally detectable, with the signal to noise ratio around $3$ or
1280: so, in a quite narrow frequency interval near and above the frequency
1281: $3\times 10^{-3} Hz$. However, at lower frequencies one would
1282: need to be concerned with the possible gravitational wave noise from
1283: unresolved binary stars in our Galaxy. The further improvement of the
1284: expected LISA sensitivity by a factor of $3$ may prove to be crucial
1285: for a confident detection of the predicted signal with $\beta = -1.9$.
1286: \par
1287: Let us now turn to the ground-based interferometers operating
1288: in the interval from $10 Hz$ to $10^{4} Hz$. The best sensitivity is
1289: reached in the band around $\nu = 10^2 Hz$. We take this frequency
1290: as the representative frequency for comparison with the predicted signal.
1291: We will work directly in terms of the dimensionless quantity $h(\nu)$.
1292: If necessary, the r.m.s. amplitude per $Hz^{1/2}$ at a given $\nu$ can be
1293: found simply
1294: as $h(\nu)/{\sqrt\nu}$. The instrumental noise will also be quoted in
1295: terms of the dimensionless quantity $h_{ex}(\nu)$.
1296: \par
1297: The expected sensitivity of the initial instruments at $\nu = 10^2 Hz$ is
1298: $h_{ex} = 10^{-21}$ or better. The theoretical prediction at this
1299: frequency, following from (\ref{47}), (\ref{48}) with $\beta_s = 0$, is
1300: $h_{th} = 10^{-23}$ for $\beta = -1.8$,
1301: and $h_{th} = 10^{-25}$ for $\beta = -1.9$. Therefore, the gap between
1302: the signal and noise levels is from 2 to 4 orders of magnitude. The expected
1303: sensitivity of the advanced interferometers, such as
1304: LIGO-II \cite{ligoII}, can be as high
1305: as $h_{ex} = 10^{-23}$. In this case, the gap vanishes for the $\beta= -1.8$
1306: signal and reduces to 2 orders of magnitude for the $\beta= -1.9$ signal.
1307: $Fig. 9$ illustrates the expected signal in comparison with the LIGO-II
1308: sensitivity. Since the signal lines are plotted in terms of $h(\nu)$, the LISA
1309: sensitivity curve (shown for periodic sources) should be raised and
1310: adjusted in accordance with $Fig. 8$.
1311:
1312:
1313:
1314:
1315: %Fig. 9
1316: \begin{figure}
1317: \epsfxsize=0.8\textwidth
1318: \centerline{\epsfbox{ligo.fig9}}
1319: \end{figure}
1320:
1321:
1322:
1323: \par
1324: A signal below noise can be detected if the outputs
1325: of two or more detectors can be cross correlated. [For the early esimates
1326: of detectability of relic gravitational waves see \cite{g4}.] The cross
1327: correlation will be possible for ground-based
1328: interferometers, several of which are currently under construction.
1329: The gap between the signal and the noise levels
1330: should be covered by a sufficiently long observation time $\tau$.
1331: The duration $\tau$ depends on whether the signal has any temporal
1332: signature known in advance, or not. We start from the assumption that no
1333: temporal signatures are known in advance. In other words, we first ignore
1334: the squeezed nature of the relic background and work under the assumption
1335: that the squeezing cannot be exploited to our advantage.
1336: \par
1337: The response of an instrument to the incoming radiation is
1338: $s(t) = F_{ij}h^{ij}$ where $F_{ij}$ depends on the position and
1339: orientation of the instrument. Since the $h^{ij}$ is a quantum-mechanical
1340: operator (see Eq. (\ref{6})) we need to calculate the mean value of a quadratic
1341: quantity. The mean value of the cross correlation of responses from two
1342: instruments $\langle 0|s_1(t)s_2(t)|0\rangle$ will involve the overlap
1343: reduction function \cite{Mich}-\cite{Allen},
1344: which we assume to be not much smaller than $1$ \cite{Flan}.
1345: The signal to noise ratio $S/N$ in the measurement of the amplitude
1346: of a signal with no specific known features increases as
1347: $(\tau \nu)^{1/4}$, where $\nu$ is some characteristic central frequency.
1348: If the signal has features known in advance and exploited by the
1349: matched filtering technique, the $S/N$ increases as
1350: $(\tau \nu)^{1/2}$.
1351: \par
1352: We apply the guaranteed law $(\tau \nu)^{1/4}$ to initial and advanced
1353: instruments at the representative frequency $\nu = 10^2 Hz$. This law
1354: requires a reasonably short time $\tau = 10^6~{\rm sec}$ in order
1355: to improve the $S/N$ in initial instruments by two orders of magnitude
1356: and to reach the level of
1357: the signal with extreme spectral index $\beta = -1.8$.
1358: The longer integration time or a better sensitivity will make the
1359: $S/N$ larger than 1. In the case of a realistic spectral index
1360: $\beta = -1.9$ the remaining gap of 4 orders of magnitude can be
1361: covered by the combination of a significantly better sensitivity and
1362: a longer observation time (not necessarily in one non-interrupted run).
1363: The sensitivity of the advanced laser interferometers, such as LIGO II,
1364: at the level $h_{ex} = 10^{-23}$ and the same observation time
1365: $\tau = 10^6~{\rm sec}$ would be sufficient for reaching the level of the
1366: predicted signal with $\beta = -1.9$.
1367: \par
1368: An additional increase of $S/N$ can be achieved if the statistical
1369: properties of the signal can be properly exploited.
1370: Squeezing is automatically present at all frequencies from $\nu_H$ to
1371: $\nu_1$. The squeeze parameter $r$ is larger in gravitational waves
1372: of cosmological scales, and possibly the periodic structure in
1373: Eq. (\ref{30}) can be better revealed at those scales. However, we
1374: are interested here
1375: in frequencies accessible to ground based interferometers,
1376: say, in the interval $30 Hz - 100 Hz$. If our intention were to monitor
1377: one given frequency $\nu$ from the beginning of its oscillating regime
1378: and up till now, then, in order to avoid the destructive interference from
1379: neighbouring modes during all that time,
1380: the frequency resolution of the instrument should have been
1381: increadibly narrow, of the order of $10^{-18} Hz$.
1382: Certainly, this is not something what we can, or
1383: intend to do. Although the amplitudes of the waves have adiabatically
1384: decreased and their frequencies redshifted since the beginning of their
1385: oscillating regime, the general statistical properties of the discussed
1386: signal are essentially the same now as they were $10$ years after the
1387: Big Bang or will be $1$ million years from now.
1388: \par
1389: The periodic structure (\ref{47}) may survive at some level in
1390: the instrumental window of sensitivity from $\nu_{min}$ (minimal frequency)
1391: to $\nu_{max}$ (maximal frequency). The mean square value of the field in
1392: this window is
1393: \begin{equation}
1394: \label{50}
1395: \int_{\nu_{min}}^{\nu_{max}}h^2(\nu,t)\frac{{\rm d}\nu}{\nu} =
1396: 10^{-14}\frac{1}{{\nu_H}^{2\beta+2}}\int_{\nu_{min}}^{\nu_{max}}
1397: \cos^2[2\pi\nu(t - t_{\nu})]\nu^{2\beta+1}{\rm d}\nu~.
1398: \end{equation}
1399: Because of the strong dependence of the integrand on frequency,
1400: $\nu^{-2.6}$ or $\nu^{-2.8}$, the value of the integral (\ref{50})
1401: is determined by
1402: its lower limit. Apparently, the search through the data should be based on
1403: the periodic structure that may survive at $\nu = \nu_{min}$. As an illustration,
1404: one can consider such a narrow interval $\Delta \nu = \nu_{max} - \nu_{min}$
1405: that the integral (\ref{50}) can be approximated by the formula
1406: \[
1407: \int_{\nu_{min}}^{\nu_{max}}h^2(\nu,t)\frac{{\rm d}\nu}{\nu} \approx
1408: 10^{-14}\left(\frac{\nu_{min}}{{\nu_H}}\right)^{2\beta+2}
1409: \left(\frac{\Delta \nu}{\nu_{min}}\right)\cos^2[2\pi\nu_{min}(t - t_{min})]~.
1410: \]
1411: Clearly, the correlation function is strictly periodic and its structure
1412: is known in advance, in contrast to other possible signals. This is a typical
1413: example of using the appriori information. Ideally, the gain in $S/N$ can
1414: grow as $(\tau \nu_{min})^{1/2}$. This would significantly reduce the
1415: required observation time $\tau$. For a larger $\Delta \nu$, even an
1416: intermediate gain between the guaranteed law $(\tau \nu)^{1/4}$ and
1417: the law $(\tau \nu)^{1/2}$, adequate for the matched filtering
1418: technique, would help. This could potentially make the signal with
1419: $\beta =-1.9$ measurable even by the initial laser interferometers.
1420: A straightforward application of (\ref{50}) for exploiting the squeezing
1421: may not be possible, as argued in the recent study \cite{AFP}, but more
1422: sophisticated methods are not excluded.
1423: \par
1424: For frequency
1425: intervals covered by bar detectors and electromagnetic detectors, the
1426: expected results follow from the same formulas (\ref{47}), (\ref{48})
1427: and have been briefly
1428: discussed elsewhere \cite{g4}, \cite{g3}.
1429:
1430:
1431: \section{Conclusion}
1432:
1433: It would be strange, if the predicted signal at the level corresponding
1434: to $\beta =-1.9$ were not seen by the instruments capable of its detection.
1435: There is not so many cosmological assumptions involved in the derivation,
1436: that could prove wrong, thus invalidating our predictions. On the other hand,
1437: it would be even more
1438: strange (and even more interesting) if the relic gravitational waves
1439: were detected at the level above the $\beta = -1.8$ line. This
1440: would mean that there is something fundamentally wrong in our
1441: basic cosmological premises. To summarise, it is quite possible that
1442: the detection of relic (squeezed) gravitational waves may be awaiting
1443: only the first generation of sensitive instruments and an appropriate
1444: data processing strategy.
1445:
1446: \section{Acknowlegements}
1447:
1448:
1449: I appreciate the help of M. V. Prokhorov in preparation of the figures.
1450:
1451:
1452: \vspace{-.5cm}
1453:
1454: \begin{thebibliography}{00}
1455:
1456: \bibitem{g1}
1457: L. P. Grishchuk, Zh. Eksp. Teor. Fiz. {\bf 67}, 825 (1974)
1458: [JETP {\bf 40}, 409 (1975)]; Ann. NY Acad. Sci. {\bf 302}, 439 (1977);
1459: Uspekhi Fiz. Nauk. {\bf 156}, 297 (1988) [Sov. Phys.-Uspekhi. {\bf 31},
1460: 940 (1988)].
1461: \bibitem{t1}
1462: K. S. Thorne, in {\it 300 Years of Gravitation}, eds. S. W. Hawking
1463: and W. Israel (Cambridge: CUP) 1987, p. 330.
1464: \bibitem{t2}
1465: K. S. Thorne, in {\it Particle and Nuclear Astrophysics and Cosmology
1466: in the Next Millenium}, eds. E. Kolb and R. Peccei (Singapore:
1467: World Scientific) 1995, p. 160.
1468: \bibitem{sch}
1469: B. F. Schutz, Class. Quant. Grav. {\bf 16}, A131 (1999).
1470: \bibitem{kn}
1471: P. L. Knight, in {\it Quantum Fluctuations}, Eds. S. Reynaud,
1472: E. Giacobino, and J. Zinn-Justin, (Elsevier Science) 1997, p. 5.
1473: \bibitem{gs}
1474: L. P. Grishchuk and Yu. V. Sidorov, Class. Quant. Grav.
1475: {\bf 6}, L161 (1989); Phys. Rev. {\bf D42}, 3413 (1990).
1476: \bibitem{g2}
1477: L. P. Grishchuk, in {\it Workshop on Squeezed States and
1478: Uncertainty Relations}, NASA Conf. Publ. {\bf 3135}, 1992, p. 329;
1479: Class. Quant. Grav. {\bf 10}, 2449 (1993);
1480: in {\it Quantum Fluctuations}, Eds. S. Reynaud,
1481: E. Giacobino, and J. Zinn-Justin, (Elsevier Science) 1997, p. 541.
1482: \bibitem{ll}
1483: L. D. Landau and E. M. Lifshitz, {\it The Classical Theory of Fields}
1484: (New York: Pergamon) 1975.
1485: \bibitem{w}
1486: W. Schleich and J. A. Wheeler, J. Opt. Soc. Am {\bf B4}, 1715 (1987);
1487: W. Schleich {\it et. al.}, Phys. Rev. {\bf A40}, 7405 (1989).
1488: \bibitem{zn}
1489: Ya. B. Zeldovich and I. D. Novikov, {\it The Structure and Evolution
1490: of the Universe} (Chicago, IL: University of Chicago Press) 1983.
1491: \bibitem{gio}
1492: M. Giovannini, Phys. Rev. {\bf D58}, 083504 (1998); Report hep-ph/9912480.
1493: \bibitem{g3}
1494: L. P. Grishchuk, Report gr-qc/9810055.
1495: \bibitem{vg}
1496: G. Veneziano, Phys. Lett. {\bf B265}, 287 (1991): M. Gasperini
1497: and G. Veneziano, Astropart. Phys. {\bf 1}, 317 (1993);
1498: M. Gasperini and M. Giovannini, Phys. Rev. {\bf D47}, 1519 (1993);
1499: M. Gasperini, Report hep-th/9607146.
1500: \bibitem{cr}
1501: T. Creighton, Report gr-qc/9907045.
1502: \bibitem{s}
1503: G. F. Smoot {\it et. al.}, Astroph. J. {\bf 396}, L1 (1992).
1504: \bibitem{b}
1505: C. L. Bennet {\it et. al.}, Astroph. J. {\bf 464}, L1 (1996).
1506: \bibitem{melch}
1507: A. Melchiori, M. V. Sazhin, V. V. Shulga, and N. Vittorio, Astroph. J.
1508: {\bf 518}, 562 (1999).
1509: \bibitem{gden}
1510: L. P. Grishchuk, Phys. Rev. {\bf D50}, 7154 (1994);
1511: in {\it Current Topics in Astrofundamental Physics: Primordial Cosmology},
1512: Eds. N. Sanchez and A. Zichichi, (Kluwer Academic) 1998, p. 539;
1513: Report gr-qc/9801011.
1514: \bibitem{g3}
1515: L. P. Grishchuk, Class. Quant. Grav. {\bf 14}, 1445 (1997).
1516: \bibitem{ms}
1517: J. Martin and D. J. Schwarz, Report astro-ph/9911225.
1518: \bibitem{Abr}
1519: A. Abramovici {\it et. al.}, Science {\bf 256}, 325 (1992).
1520: \bibitem{Brad}
1521: C. Bradaschia {\it et. al.}, Nucl. Instrum. and Methods
1522: {\bf A289}, 518 (1990).
1523: \bibitem{HoughD}
1524: J. Hough and K. Danzmann {\it et. al.}, {\it GEO600 Proposal}, 1994.
1525: \bibitem{lisa}
1526: P. Bender {\it et. al.} {\it LISA Pre-Phase A Report}, Second Edition,
1527: 1998
1528: \bibitem{Hel}
1529: S. L. Larson, W. A. Hiscock. and R. W. Hellings, Report gr-qc/9909080.
1530: \bibitem{gw0}
1531: {\it Gravitational Wave Experiments}, eds. E. Coccia, G. Pizzella, and
1532: F. Ronga (Singapore: World Scientific) 1995.
1533: \bibitem{gw1}
1534: {\it Gravitational Waves: Sources and detection}, eds. I. Giufolini
1535: and F. Fidecaro (Singapore: World Scientific) 1997.
1536: \bibitem{gw2}
1537: {\it Laser Interferometer Space Antenna}, ed. W. Folkner (AIP Conference
1538: Proceedings 456) 1998.
1539: \bibitem{ligoII}
1540: {\it LIGO II Conceptual Project Book}, LIGO-M990288-00-M.
1541: \bibitem{g4}
1542: L. P. Grishchuk, Pis'ma Zh. Eks. Teor. Fiz. {\bf 23}, 326 (1976) [JETP
1543: Lett. {\bf 23}, 293 (1976)].
1544: \bibitem{Mich}
1545: P. F. Michelson, Mon. Not. R. Astr. Soc. {\bf 227}, 933 (1987).
1546: \bibitem{Chris}
1547: N. L. Christensen, Phys. Rev. {\bf D46}, 5250 (1992).
1548: \bibitem{Flan}
1549: E. E. Flanagan, Phys. Rev. {\bf D48}, 2389 (1993).
1550: \bibitem{Allen}
1551: B. Allen, in {\it Relativistic Gravitation and Gravitational
1552: Radiation}, Eds. J-A. Marck and J-P. Lasota (CUP, 1997) p. 373.
1553: \bibitem{AFP}
1554: B. Allen, E. E. Flanagan, and M. A. Papa. Report gr-qc/9906054.
1555:
1556:
1557: \end{thebibliography}
1558: \end{document}
1559:
1560:
1561:
1562: