1: %%%\documentstyle[preprint,aps,floats,epsfig,subeqn,amssymb]{revtex}
2:
3: \documentstyle[preprint,aps,epsfig,floats]{revtex}
4:
5:
6:
7:
8:
9:
10:
11: % Uncomment next two lines for A4 paper size, comment for letter size
12:
13: %\addtolength{\textheight}{17.6mm}
14:
15:
16:
17: %%%%% number equations by section %%%%%%%%
18:
19: %\makeatletter
20:
21: %\@addtoreset{equation}{section}
22:
23: %\makeatother
24:
25:
26:
27: % Command Definitions
28:
29:
30:
31: \def\real{I\negthinspace R}
32:
33: \def\zed{Z\hskip -3mm Z }
34:
35: \def\half{\textstyle{1\over2}}
36:
37: \def\quarter{\textstyle{1\over4}}
38:
39: \def\sech{\,{\rm sech}\,}
40:
41: \def\ie{{\it i.e.,}}
42:
43: \newcommand{\be}{\begin{equation}}
44:
45: \newcommand{\ee}{\end{equation}}
46:
47: \newcommand{\bea}{\begin{eqnarray}}
48:
49: \newcommand{\eea}{\end{eqnarray}}
50:
51: \newcommand{\bml}{\begin{mathletters}}
52:
53: \newcommand{\eml}{\end{mathletters}}
54:
55: %\newcommand{\bml}{\begin{subequations}}
56:
57: %\newcommand{\eml}{\end{subequations}}
58:
59: %
60:
61: \def\aprle{\buildrel < \over {_{\sim}}}
62:
63: \def\aprge{\buildrel > \over {_{\sim}}}
64:
65:
66:
67: \begin{document}
68:
69:
70:
71: \tighten
72:
73:
74:
75: \preprint{DCPT-03/07}
76:
77: \draft
78:
79:
80:
81: %\twocolumn[\hsize\textwidth\columnwidth\hsize\csname @twocolumnfalse\endcsname
82:
83:
84:
85: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
86:
87:
88:
89: %\wideabs{ % Uncomment this line for two-column output
90:
91:
92:
93: \title{Strings in de Sitter space }
94:
95: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
96:
97:
98:
99: \author{Eug\^enio R. Bezerra de Mello\footnote{emello@fisica.ufpb.br}}
100: \address{Departamento de F\'{\i}sica-CCEN, Universidade Federal da
101: Para\'{\i}ba, 58.059-970, J. Pessoa, PB, C. Postal 5.008, Brazil}
102: \author{ Yves Brihaye\footnote{Yves.Brihaye@umh.ac.be}}
103: \address{Facult\'e des Sciences, Universit\'e de Mons-Hainaut,
104: B-7000 Mons, Belgium}
105: \author{Betti Hartmann\footnote{Betti.Hartmann@durham.ac.uk}}
106: \address{Department of Mathematical Sciences, University
107: of Durham, Durham DH1 3LE, U.K.}
108: \date{\today}
109:
110: \setlength{\footnotesep}{0.5\footnotesep}
111:
112:
113:
114: \maketitle
115:
116: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
117:
118: \begin{abstract}
119:
120: We study both global as well as local (Nielsen-Olesen) strings
121: in de Sitter space. While these type of topological defects
122: have been studied in the background of a de Sitter metric previously, we study
123: here the full set of coupled equations.
124: We find only ``closed'' solutions. The behaviour of the metric tensor
125: of these solutions resembles that of ``supermassive'' strings with a curvature
126: singularity at the cosmological horizon.
127: For global strings (and the composite defect) we are able to construct
128: solutions which are regular on the interval from the origin to
129: the cosmological horizon if the global string core lies completely inside the
130: horizon.
131: \end{abstract}
132:
133:
134:
135: \pacs{PACS numbers: 04.20.Jb, 04.40.Nr, 11.27.+d }
136:
137:
138:
139: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
140:
141: %\newpage
142:
143: \renewcommand{\thefootnote}{\arabic{footnote}}
144:
145: \section{Introduction}
146: A number of different topological defects \cite{shell} are thought to have been
147: formed during the phase transitions in the early universe. Depending on the topology
148: of the vacuum manifold ${\cal M}$ these are domain walls, strings, monopoles and textures corresponding
149: to the homotopy groups $\pi_0({\cal M})$, $\pi_1({\cal M})$, $\pi_2({\cal M})$ and $\pi_3({\cal M})$, respectively.
150: Cosmic strings \cite{shell,kibble} have always gained a lot of interest since they are thought
151: to be important for the structure formation in the universe due to their huge
152: energy per unit length (roughly $10^{21}\frac{kg}{m}$ for a string formed at
153: GUT scale $\approx 10^{16} GeV$).
154:
155: A classical field theory model which has string-like solutions is the Abelian Higgs model \cite{no}. These solutions, also sometimes called ``vortices'', correspond to
156: infinitely long objects. They have a core radius inverse proportional
157: to the Higgs boson mass and magnetic flux tubes with radius proportional to the
158: inverse of the gauge boson mass. Coupling the Abelian Higgs model minimally to gravity, the influence of the vortex on the geometry of space-time was investigated analytically \cite{Gar}. It was shown that far away from the core
159: of the string the space-time is Minkowski minus a wedge. It was
160: also realised \cite{laguna} that if the vacuum expectation value of
161: the Higgs field is sufficiently large (corresponding
162: to strings having formed at a phase transition with energy scale much higher
163: than the GUT scale), then a different type of solution
164: is possible. These so-called ``supermassive strings'' exist only on a
165: finite interval of the radial coordinate and have a curvature singularity at
166: the maximal value of the radial coordinate.
167: The existence of further solutions was investigated in a detailed numerical
168: analysis \cite{Christensen,yves}. It was found that the parameter space
169: is indeed divided by the curve of maximal angular deficit $2\pi$.
170: If the deficit is smaller than $2\pi$, so-called ``open'',
171: i.e. infinitely extended solutions were found. One is the above mentioned
172: cosmic string solution \cite{Gar} which however has a ``shadow'' solution
173: of Melvin-type for all values of the coupling constants.
174: For deficit large than $2\pi$, only the ``supermassive'',
175: ``closed'' solutions exist \cite{laguna}.
176:
177: The static solutions of the model without gauge field, so-called global
178: strings, have also been studied \cite{cohen,harari,Gibbons,gregory1}. Like all global defects, the global string has a
179: long-range Goldstone field which leads to a divergent energy.
180: Moreover, the global string is characterised by a logarithmically
181: divergent deficit angle in contrast to the local
182: string which has a constant deficit angle. The coupling to gravity in the case
183: of the static global monopole leads to a singularity-free monopole solution
184: \cite{vilenkin} in the sense that while the energy is still
185: linearly divergent, the solid deficit angle is now finite. For
186: the static global string the corresponding singularity can not be removed
187: by coupling the system
188: to gravity and only the assumption that the metric be time-dependent
189: removes the singularity \cite{gregory2}.
190:
191: Since a number of astrophysical observations like e.g. the measurement of
192: redshifts of Type Ia supernovae \cite{super} has led scientists to believe that we live in a universe with positive cosmological constant, the study of topological
193: defects in de Sitter (dS) space seems interesting.
194: But it also is of interest from another point of view, namely the dS/CFT
195: correspondence \cite{strominger}. This correspondence
196: suggests a holographic duality between gravity in a $d$-dimensional
197: dS space and a conformal field theory (CFT) ``living''
198: on the boundary of the dS spacetime and thus being $d-1$-dimensional.
199:
200: Recently, Nielsen-Olesen strings in the background
201: of a $4$-dimensional de Sitter spacetime $dS_4$ have been studied \cite{GM}.
202: However, to our knowledge, the {\it full} system of coupled matter and metric
203: field equations has not been studied yet. One of the aims of this
204: paper is the investigation of exactly this point.
205:
206: Motivated by some recent work on
207: a composite system of a
208: global and local monopole in curved space-time \cite{bbh,by,spi},
209: we investigate the composite system of a global and
210: Nielsen-Olesen string in a curved space-time with cosmological constant
211: as well.
212:
213: Our paper is organised as follows: we give the model and static, cylindrically
214: symmetric Ansatz in Section 2. We give the equations of motion in
215: Section 3. We discuss the pure Nielsen-Olesen solutions
216: in Section 4, the global string solutions
217: in Section 5 and the composite system of a global and Nielsen-Olesen
218: string in Section 6. We give our summary in Section 7.
219:
220:
221: \section{The Model}
222: %%%%%%%%%%%%%%%%%%%%%
223: The model which describes a gravitating Nielsen-Olesen string interacting with
224: a global one in the presence of a non-vanishing cosmological constant is
225: given by the following action:
226:
227: \begin{equation}
228: \label{action}
229: S=\int d^4 x \sqrt{-g} \left( \frac{1}{16\pi G} (R-2\Lambda) + {\cal L}_{NO}+
230: {\cal L}_{global} + {\cal L}_{inter} \right)
231: \end{equation}
232: where $R$ is the Ricci scalar, $G$ denotes Newton's constant and $\Lambda$ is the
233: cosmological constant. The Lagrangian of the Abelian Higgs model is given by
234: \cite{no}:
235: \begin{equation}
236: {\cal L}_{NO}=\frac{1}{2}D_{\mu} \phi (D^{\mu} \phi)^*-\frac{1}{4} F_{\mu\nu} F^{\mu\nu}
237: -\frac{\lambda_1}{4}\left(\phi\phi^*-\eta^2_1\right)^2
238: \end{equation}
239: with the covariant derivative $D_\mu=\nabla_{\mu}-ieA_{\mu}$ and the
240: field strength $F_{\mu\nu}=\partial_\mu A_\nu-\partial_\nu A_\mu$ of the U(1) gauge potential $A_{\mu}$ with coupling constant $e$.
241: $\phi$ is a complex scalar field (the Higgs field) with vacuum expectation value $\eta_1$ and self-coupling constant $\lambda_1$.
242: The Lagrangian of the global string reads \cite{shell}:
243: \begin{equation}
244: {\cal L}_{global}=\frac{1}{2}\partial_{\mu} \chi \partial^{\mu} \chi^* - \frac{\lambda_2}{4}\left(\chi\chi^*-\eta^2_2\right)^2
245: \end{equation}
246: where $\chi$ is a complex scalar field (the Goldstone field) with vacuum expectation value $\eta_2$ and self-coupling $\lambda_2$.
247: Finally, following \cite{bbh} we introduce an extra potential which couples (with coupling constant $\lambda_3$)
248: the two sectors of the model directly to each other:
249: \begin{equation}
250: {\cal L}_{inter}=-\frac{\lambda_3}{4}\left(\phi\phi^*-\eta^2_1\right)\left(\chi\chi^*-\eta^2_2\right)
251: \end{equation}
252: Without this term, the global and local string would be coupled only indirectly
253: over gravity. In this paper we will use units which $\hbar=c=1.$
254:
255:
256:
257: \subsection{The Ansatz}
258:
259: In the following we shall analyse the classical equations of motion associated
260: with the above system. In order to do that,
261: let us write down the matter and gravitational fields as shown below.
262: The most general, cylindrically symmetric line element invariant under boosts
263: along the $z-$direction is:
264: \begin{equation}
265: ds^2=N^2(\rho)dt^2-d\rho^2-L^2(\rho)d\varphi^2-N^2(\rho)dz^2 \ .
266: \end{equation}
267: The non-vanishing components of the Einstein tensor $G_{\mu\nu}$ then read:
268: \begin{eqnarray}
269: & & G_{tt}=-G_{zz}=\frac{N}{L}\left(L\partial_{\rho\rho}N + \partial_{\rho}N \partial_{\rho}L+
270: N\partial_{\rho\rho}L \right), \nonumber \\
271: & & G_{rr}=\frac{\partial_{\rho}N}{N^2L}\left(2\partial_{\rho}L N+\partial_{\rho}N L\right)\ \ ,
272: \ \ G_{\varphi\varphi}=\frac{L^2}{N^2}\left(2N\partial_{\rho\rho}N+(\partial_{\rho} N)^2\right) \ ,
273: \end{eqnarray}
274: where $\partial_{\rho}$ denotes the derivative with respect to $\rho$.
275:
276: For the matter and gauge fields, we have:
277: \begin{equation}
278: \phi(\rho,\varphi)=\eta_1 h(\rho)e^{i n\varphi} \ ,
279: \end{equation}
280: \begin{equation}
281: \chi(\rho,\varphi)=\eta_1 f(\rho)e^{i m\varphi} \ ,
282: \end{equation}
283: \begin{equation}
284: A_{\mu}dx^{\mu}=\frac {1}{e}(n-P(\rho)) d\varphi \ .
285: \end{equation}
286: $n$ and $m$ are integers indexing the vorticity of the Higgs and Goldstone fields, respectively, around the $z-$axis.
287:
288: Substituting the above configurations into the matter Lagrangian density
289: ${\cal L}_{M}={\cal L}_{NO}+ {\cal L}_{global} + {\cal L}_{inter}$, we obtain:
290: \begin{eqnarray}
291: {\cal L}_M&=&-\frac{\eta_1^2}2(\partial_{\rho}h(\rho))^2-
292: \frac{\eta_1^2}2(\partial_{\rho}f(\rho))^2-\frac{n^2}
293: {2e^2L^2(\rho)}(\partial_{\rho}P(\rho))^2\nonumber\\
294: &-&\frac{\eta_1^2n^2}{2L^2(\rho)}h^2(\rho)P^2(\rho)-
295: \frac{\eta_1^2m^2}{2L^2(\rho)}f^2(\rho)-
296: \frac{\lambda_1\eta_1^4}4(h^2(\rho)-1)^2\nonumber\\
297: &-&\frac{\lambda_2\eta_1^4}4(f^2(\rho)-q^2)^2
298: -\frac{\lambda_3\eta_1^4}2 (h^2(\rho)-1)(f^2(\rho)-q^2) \ .
299: \end{eqnarray}
300:
301: \section{Equations of Motion}
302:
303: We define the following dimensionless variable and function:
304: \begin{equation}
305: x=\sqrt{\lambda_1}\eta_1 \rho \ \ \ , \ \ \ L(x)=L(\rho)\eta_1\sqrt{\lambda_1} \ .
306: \end{equation}
307: Then, the total Lagrangian only depends on the following dimensionless coupling constants
308: \begin{equation}
309: \gamma=8\pi G\eta_1^2 \ \ , \ \alpha=e^2/\lambda_1 \ \ , \ \
310: q=\frac{\eta_2}{\eta_1} \ \ , \ \ {\bar\Lambda}=
311: \frac\Lambda{\lambda_1\eta_1^2} \ \ , \ \
312: \beta_i^2=\frac{\lambda_i}{\lambda_1} \ \ , \ \ i=1,2,3 \ .
313: \end{equation}
314:
315: Varying (\ref{action}) with respect to the matter fields and metric functions, we obtain a system of five non-linear differential equations. The Euler-Lagrange equations for the matter field functions read:
316: \begin{equation}
317: \frac{(N^2Lh')'}{N^2L}=\frac{n^2}{L^2} hP^2+h(h^2-1)+\beta_3^2h(f^2-q^2) \ ,
318: \end{equation}
319: \begin{equation}
320: \frac{(N^2Lf')'}{N^2L}=\frac{m^2f}{L^2}+\beta_2^2f(f^2-q^2)+\beta_3^2
321: f(h^2-1) \ ,
322: \label{eqf}
323: \end{equation}
324: \begin{equation}
325: \frac{L}{N^2}\left(\frac{N^2P'}{L}\right)'=\alpha h^2P \ ,
326: \end{equation}
327: while the Einstein equations
328: \begin{equation}
329: G_{\mu\nu}+\bar\Lambda g_{\mu\nu}=\gamma T_{\mu\nu} \ \ , \ \ \mu,\nu=t,x,\varphi,z
330: \end{equation}
331: read:
332: \begin{eqnarray}
333: \label{N1}
334: \frac{(LNN')'}{N^2 L}&=&-{\bar\Lambda}+\gamma\left[\frac{n^2(P'(x))^2}
335: {2\alpha L^2}-\frac14(h^2(x)-1)^2-\frac{\beta_2^2}4(f^2(x)-q^2)^2\right.
336: \nonumber\\
337: &-&\left.\frac{\beta_3^2}2
338: (h^2(x)-1)(f^2(x)-q^2)\right]
339: \end{eqnarray}
340: and
341: \begin{eqnarray}
342: \label{N2}
343: \frac{(N^2L')'}{N^2L}&=&-{\bar\Lambda}-\gamma\left[\frac{n^2h^2(x)P^2(x)}
344: {L^2(x)}+\frac{m^2f^2(x)}{L^2(x)}+\frac{n^2(P'(x))^2}{2\alpha L^2(x)}+
345: \frac14(h^2(x)-1)^2
346: \right.\nonumber\\
347: &+&\left.\frac{\beta_2^2}4(f^2(x)-q^2)^2+\frac{\beta_3^2}2
348: (h^2(x)-1)(f^2(x)-q^2)\right]
349: \end{eqnarray}
350: Moreover, defining $u=\sqrt{-g}=N^2L$ we get the following equation:
351: \begin{eqnarray}
352: \label{ueq}
353: \frac{u''(x)}{u(x)}&=&-3{\bar\Lambda}-\gamma\left[\frac{n^2h^2(x)P^2(x)}
354: {L^2(x)}+\frac{m^2f^2(x)}{L^2(x)}-\frac{n^2(P'(x))^2}{2\alpha L^2(x)}\right.
355: \nonumber\\
356: &+&\frac34(h^2(x)-1)^2+\frac{3\beta_2^2}4(f^2(x)-q^2)^2\nonumber\\
357: &+&\left.\frac{3\beta_3^2}2(h^2(x)-1)(f^2(x)-q^2)\right] \ .
358: \end{eqnarray}
359: The prime now denotes the derivatives with respect to $x$.
360:
361: \subsection{Boundary Conditions}
362:
363: The requirement of regularity at the origin leads to the following boundary
364: conditions:
365: \begin{equation}
366: h(0)=0, \ f(0)=0 \ , \ P(0)=n \
367: \end{equation}
368: for the matter fields and
369: \begin{equation}
370: \label{zero}
371: N(0)=1, \ N'(0)=0, \ L(0)=0 \ , \ L'(0)=1 \ .
372: \end{equation}
373: for the metric fields. Since a cosmological horizon appears naturally in de Sitter
374: space, we integrate the equations only up to this value of the coordinate $x$,
375: $x=x_0$. In order for the core of the local string to lie
376: completely within the horizon we require:
377: \begin{equation}
378: h(x=x_0)=1, \ f(x=x_0)=q \ , \ P(x=x_0)=0 \ .
379: \end{equation}
380: Note that due to the fact that the $x$ interval is finite, this is not (like in asymptotically
381: flat space) a necessary condition for finite energy solutions. However, we have chosen
382: these boundary conditions such that the energy-momentum tensor vanishes at $x=x_0$.
383: In addition, the limit $\bar{\Lambda}\rightarrow 0$ which leads to $x_0\rightarrow \infty$
384: can be taken with these boundary conditions.
385:
386: \section{Nielsen-Olesen strings in de Sitter space}
387:
388: First, we are interested in the case of the pure Nielsen-Olesen string.
389: This corresponds to setting $f(x)\equiv 0$ and $q\equiv 0$ in the previous equations.
390: \subsection{Vacuum solution}
391:
392: For the case of the pure gauge string, there is a vacuum solution of the equations.
393: Setting $P(x)=0$ and $h(x)=1$, we find from (\ref{ueq}), that:
394: \begin{equation}
395: N^2(x)L(x)=A\sin(\sqrt{3\bar\Lambda}x)+B\cos(\sqrt{3\bar\Lambda}x) \ , \ \ A, B \ \ {\rm constants} \ .
396: \end{equation}
397: Using the boundary conditions (\ref{zero}), we find the following solution:
398: \begin{equation}
399: N^2(x)L(x)=\frac{1}{\sqrt{3\bar\Lambda}}\sin(\sqrt{3\bar\Lambda}x) \ .
400: \end{equation}
401: This then can be put into (\ref{N1}) and (\ref{N2}) and we find the solutions:
402: \begin{equation}
403: N(x)=\cos^{2/3}(\sqrt{3\bar\Lambda}\frac x2)
404: \label{NON}
405: \end{equation}
406: and
407: \begin{equation}
408: L(x)=\frac{2^{2/3}}{\sqrt{3\bar\Lambda}}
409: [\sin(\sqrt{3\bar\Lambda} x)]^{1/3}[\tan(\sqrt{3\bar\Lambda}
410: \frac{x}{2})]^{2/3}
411: \label{NOL}
412: \end{equation}
413: where the coefficients again result from the
414: boundary conditions (\ref{zero}). The first zero of $N(x)$ lies at
415: $x^v_0=\pi/\sqrt{3\bar\Lambda}$. At the same time,
416: $L(x\rightarrow x^v_0)\rightarrow \infty$. This is the cosmological horizon
417: of the vacuum solution. If we expand the metric functions around this horizon, we obtain:
418: \begin{equation}
419: N(x\rightarrow x^v_0)\approx(-\frac{\sqrt{3\bar\Lambda}}{2})^{2/3} (x-x^v_0)^{2/3}+...
420: \end{equation}
421: and
422: \begin{equation}
423: L(x\rightarrow x^v_0)\approx(\frac{\sqrt{3\bar\Lambda}}{2})^{-4/3} (x-x^v_0)^{-1/3}+.. \ .
424: \end{equation}
425: This has the behaviour of a so-called Kasner solution \cite{kramer}:
426: \begin{equation}
427: ds^2=(k\rho)^{2a}dt^2-d\rho^2-C^2(k\rho)^{2(b-1)}\rho^2 d\varphi^2-(k\rho)^{2c}dz^2
428: \end{equation}
429: with $a=c=2/3$, $b=-1/3$, $k=\frac{\sqrt{3\bar\Lambda}}{2}$, $C=1$.
430: These type of ``closed'' solutions have been found previously \cite{laguna} for the case $\bar\Lambda=0$
431: and were called ``supermassive'' strings. When calculating
432: the Kretschmann scalar $K=R^{\mu\nu\rho\sigma}R_{\mu\nu\rho\sigma}$ one obtains \cite{laguna} that $K\propto (x-x^v_0)^{-4}$ and thus the solution has indeed a curvature singularity at $x=x^v_0$. Remarkable is that in the case of $\bar\Lambda=0$, these type of solutions only appear for
433: sufficiently high vacuum expectation values (vev) of the Higgs field corresponding
434: to strings having formed at energy scales much higher than the
435: GUT scale \cite{laguna}. For smaller values of the vev no singularity appears
436: and the solutions exist on the full interval $[0:\infty[$.
437: Accordingly, the numerical study showed \cite{yves} that these solutions
438: exist for a $\gamma > \gamma_{cr}$.
439:
440:
441: \subsection{Numerical results}
442:
443: Subject to the boundary conditions (\ref{zero}), we have studied the coupled
444: system of equations numerically.
445:
446: First, we fixed $\alpha$ and $\bar\Lambda$ to study the influence of
447: the gravitational coupling $\gamma$ on the solutions.
448: We determined the value of $x$ at which the metric function $N(x)$ vanishes, i.e. $N(x=x_0)=0$. Our results for $\alpha=1.0$, $\bar\Lambda=0.005$ and $n=1$, $2$ are shown in Fig.~1. As expected the value of $x_0$ decreases with increasing gravitational coupling $\gamma$. Moreover, we observe a steep decrease in $x_0$ for a relatively small range of $\gamma$. We have only plotted results for $\gamma$'s corresponding to $x_0 \geq 5$ since for large $\gamma$'s the numerics
449: becomes increasingly difficult. The reason for this is indicated in Fig.~2, where we
450: show the profiles of the metric functions $N(x)$, $L(x)$ as well as those of the matter field functions $P(x)$ and $h(x)$ for $\alpha=1.0$, $\bar\Lambda=0.005$, $n=1$
451: and two different
452: choices of $\gamma$. For $\gamma=1.5$, the value of $x$ at which the matter field functions reach their asymptotic values $0$ and $1$, respectively, is much smaller than the value of $x_0$.
453: This means that the horizon clearly lies outside the core of the string. For $\gamma=1.7$, however, the situation is different. The value of $x$ at which $h(x)$ reaches $1$ is roughly equal to $x_0$, while $P(x)$ seems to be still greater than $0$ on the plot we present. This is due to the fact that the ``real'' solution would have a slightly higher $x_0$ at which $P(x=x_0)=0$. However, since
454: $L(x\rightarrow x_0)\rightarrow \infty$, it is numerically impossible to reach the final solution. Nevertheless, the plot indicates that -like $h(x)$- $P(x)$ just reaches its asymptotic value $0$ at $x=x_0$. Thus the horizon lies very close to the core of the string.
455:
456:
457: Then, we fixed $\alpha$ and $\gamma$ and determined $x_0$ in dependence on $\bar\Lambda$. Our results for $\alpha=\gamma=0.5$ and $n=1$, $2$ together with the location
458: of the cosmological horizon of the vacuum solution, $x^v_0$, are given in Fig.~3.
459:
460: We clearly observe that the value of the cosmological horizon
461: decreases with the increase of the cosmological constant, as expected.
462: Moreover, an increase in the vorticity $n$ leads to a decrease of $x_0$ for
463: the same $\bar\Lambda$. For small $\bar\Lambda$, $x_0$ of both the
464: $n=1$ and the $n=2$ solution is very close to the corresponding $x^v_0$.
465: This can again be explained by studying the behaviour of the
466: matter field functions for varying $\bar{\Lambda}$.
467: As observed previously \cite{GM}, we find that for fixed $\alpha$ and $\gamma$
468: and increasing $\bar{\Lambda}$, the value of the coordinate $x$
469: at which the matter field functions reach their asymptotic values
470: also increases, e.g. for $\alpha=\gamma=0.5$, $n=2$ we find
471: that for $\bar{\Lambda}=0.001$ the value of $x$, where the gauge field function
472: reaches $P(x^{0.1})=0.1$ is $x^{0.1}(\bar{\Lambda}=0.001)\approx 6.2$,
473: while for $\bar{\Lambda}=0.005$, we find
474: $x^{0.1}(\bar{\Lambda}=0.005)\approx 6.45$. This can be interpreted
475: as representing a thicker string core due to an increased cosmological expansion.
476: Thus the cosmological horizon lies closer and closer to the core of the string
477: for increasing $\bar\Lambda$ and so only for small $\bar\Lambda$
478: the solution close to the cosmological horizon can be described by the
479: vacuum solution. In Fig.~4, we show the profiles of $N(x)$ and $L(x)$
480: for $\bar\Lambda=0.01$, $\alpha=\gamma=0.5$ and $n=2$ both for the Nielsen-Olesen string and the vacuum solution. Clearly, the solutions differ quite strongly.
481:
482: \section{Global strings in de Sitter space}
483:
484:
485:
486: This is the case of setting $P(x)\equiv n$ and $h(x)\equiv 0$ (which implies $\eta_1\equiv 0$). The global string without a cosmological constant, i.e. $\bar\Lambda=0$ has been studied extensively \cite{cohen,harari,Gibbons,gregory1}. To have a good starting solution for the construction of dS global strings, we have reconstructed the global string without cosmological constant. We find that the metric function $N(x)$ deviates very little from one, but that the quantity $|N(x=0)-N(x^*)|$ is increasing
487: with $x^*\rightarrow\infty$. Moreover, $L(x)$ grows approximately linearly with $x$.
488: This is the behaviour found in \cite{harari}, namely that outside the core of the global string the metric functions behave like
489: $N^2=1-\frac{\gamma}{2}\ln(\frac{x}{x_{gc}})$, $L^2=x^2(1-\gamma\ln(\frac{x}{x_{gc}}))$ with $x_{gc}\propto (q\beta_2)^{-1}$ being the core of the global string in our rescaled coordinates. Moreover, we recover the behaviour of the Goldstone field function $f(x\rightarrow \infty)=q-O(x^{-2})$.
490:
491: \subsection{Numerical results}
492:
493:
494:
495: For non-vanishing cosmological constant we find that the behaviour of $N(x)$ and
496: $L(x)$ resembles that of the metric functions in the case of the Nielsen-Olesen
497: string in de Sitter space (see previous section). Again, for all
498: constants fixed and $\bar\Lambda$ varied, we find that
499: the value at which $N(x=x_0)=0$ decreases with increasing
500: $\bar\Lambda$, e.g. for $\beta_2=1$, $q=0.1$, $\gamma=0.1$, we find that
501: $x_0(\bar\Lambda=10^{-4})\approx 180$, while $x_0(\bar\Lambda=10^{-3})\approx 57$.
502: At the same time, the value of $x$ at which the function $f(x)$ reaches its
503: vev increases with increasing $\bar\Lambda$ which is due to
504: the increased cosmological expansion thus leading to an extended string core.
505: The behaviour of the function $f(x)$ depends crucially on the
506: cosmological constant which determines the cosmological horizon
507: and the parameters $\beta_2$ and $q$, which determine the radius of the string core.
508: We will discuss these features in more detail in the context of composite topological
509: defects in the next section.
510:
511:
512:
513: \section{Composite system of global and Nielsen-Olesen string}
514: Unlike in the case of the ``pure'' Nielsen-Olesen string, a complete analytical analysis
515: of the composite system seems to be impossible.
516: However, some additional information can be gained by analysing the energy density per unit
517: length. Before discussing our numerical results, we make some remarks on this point in the
518: following section.
519:
520:
521:
522: The energy density per unit length of the composite defect is given by \cite{Christensen,yves}:
523: \begin{equation}
524: {\cal E}=\int \sqrt{-g_3} T^0_0 dx_1 dx_2
525: \end{equation}
526: where $g_3$ is the determinant of the $2+1$ dimensional metric
527: $ds^2=N^2(\rho)dt^2-d\rho^2-L^2(\rho)d\varphi^2$ and $T^0_0=-{\cal L}_M$ is the
528: $00$-component of the energy-momentum tensor. In the cylindrical coordinates,
529: we get:
530: \begin{eqnarray}
531: \label{E}
532: {\cal E}&=&\pi\eta_1^2\int_0^{x_0}d x N(x)L(x)\left[
533: \frac{n^2}{\alpha L^2(x)}(P'(x))^2+(h'(x))^2+(f'(x))^2\right.\nonumber\\
534: &+&\frac{n^2h^2(x)P^2(x)}{L^2(x)}+\frac{m^2f^2(x)}{L^2(x)}+
535: \frac12(h^2(x)-1)^2+\frac{\beta_2^2}2(f^2(x)-q^2)^2\nonumber\\
536: &+&\left.\beta_3^2(h(x)^2-1)(f^2(x)-q^2)\right] \ .
537: \end{eqnarray}
538: From this, we can observe that the composite defect
539: has a finite energy density. There are two different scenarios now. If the product
540: $\beta_2 q$ is large enough the core of the global string is small and
541: the cores of both strings lie within the cosmological horizon $x_0$. We can then
542: assume that $f(x_0)=q$, $f^{'}(x_0)=0$, $P(x_0)=0$, $h(x_0)=1$.
543: From before, we know that $L(x)$ goes to infinity close to the horizon, while
544: $N(x)L(x)$ remains finite. Thus the integrand of (\ref{E}) tends to zero like
545: $m^2q^2\frac{N(x)}{L(x)}$. If the product $\beta_2 q$ is small, the
546: core of the global string extends to outside the horizon $x_0$.
547: Then the integrand of (\ref{E}) becomes
548: $N(x)L(x)\left[(f'(x))^2+\frac{m^2f^2(x)}{L^2(x)}+\frac{\beta_2^2}2(f^2(x)-q^2)^2\right]$.
549: Since close to the horizon, the product $N(x)L(x)$ tends to
550: zero, this is finite. We have indeed confirmed numerically that this is the case.
551:
552:
553: \subsection{Numerical Results}
554:
555: Assuming that the behaviour of the metric functions $N(x)$ and $L(x)$ persists in the presence
556: of all (non-trivial) matter fields (which indeed our numerical analysis confirms),
557: we can insert (\ref{NON}) and (\ref{NOL}) into (\ref{eqf}). We obtain that:
558: \begin{equation}
559: f(x\rightarrow x_0)\sim q+ C(x-x_0)^{8/3} \ \ , \ \ \ C \ \ {\rm constant}
560: \end{equation}
561:
562: We have construct the composite model solution
563: by starting from the Nielsen-Olesen (NO) string and increasing the parameter
564: $q$ gradually from zero.
565: Our numerical analysis confirms the assumption that the radius of the NO
566: string core is smaller than that of the global string. Moreover, we find
567: that the NO string always resides inside
568: the horizon.
569: As for the global string, we find that when the coupling constant $\beta_2$ is small
570: the function $f(x)$
571: reaches the imposed expectation value $f(x_0) = q$ with a positive
572: concavity, in particular the derivative $f'|_{x=x_0}$ is non zero
573: and our numerical solution clearly suggests that the behaviour
574: $f(x\rightarrow x_0) \sim q + C (x - x_0)^{8/3}$ is not fulfilled.
575: This is demonstrated in Fig.~5, where we show $f(x)$ of the composite
576: defect for $\beta_2^2=1$, $\alpha=\gamma=\beta_1^2=1$, $q=0.1$, $\beta_3^2=0$ and
577: $\bar\Lambda=0.001$.
578: This suggests that for small values of the product $\beta_2 q$, the argument demonstrated
579: above doesn't hold.
580:
581: Increasing the value of $\beta_2$ we were able to
582: produce solutions which seem to have the expected behaviour, i.e.
583: $f(0) =0$, $f(x=x_0)=q$ and $f'|_{x=x_0}=0$. This is shown in Fig.~5
584: for $\alpha=\gamma=\beta_1^2=1$, $q=0.1$, $\beta_3^2=0$,
585: $\bar\Lambda=0.001$ and $\beta_2^2=5$, $10$, $20$, $50$, respectively.
586: Of course, the occurrence
587: of the singularity at $x = x_0$ renders the interpretation of the numerical results
588: not hundred percent certain but we are rather confident that
589: composite string defects which are regular inside the horizon
590: exist for large enough $\beta_2$.
591:
592: All these results are obtained for the case of $\beta_3=0$, i.e. the two defects
593: interact with each other only indirectly over gravity.
594: We have also attempted to construct solutions with $\beta_3\neq 0$.
595: We find that only for large enough values of the quotient $\beta_2^2/\beta_3^2$,
596: the solutions seem well behaved. If $\beta_2^2/\beta_3^2$ is roughly of the order
597: of $10^2$ (for $\alpha=\gamma=\beta_1^2=1$, $q=0.1$ and $\bar\Lambda=0.001$),
598: the behaviour of the functions is
599: very similar to that in the case of $\beta_3=0$. For $\beta_2^2/\beta_3^2$
600: smaller than that, however, the functions $f(x)$ and $h(x)$ start to develop
601: oscillations close to the cosmological horizon. The number of oscillations
602: increases with the decrease of the quotient $\beta_2^2/\beta_3^2$.
603: Thus, we conclude that directly interacting
604: composite defects without the global string singularity only
605: exist if the self-interaction of the global string
606: is much larger than the interaction between the two defects.
607:
608:
609:
610: \section{Conclusion}
611:
612: In this paper we have analysed both Nielsen-Olesen and global strings as
613: well as the composite system of both defects in de Sitter space.
614: When the matter fields are set equal to their vacuum expectation
615: values (vev) in the case of the ``pure'' Nielsen-Olesen string,
616: we were able to construct analytic solutions of the Einstein equations
617: in terms of trigonometric functions. The metric tensor resembles that
618: of so-called ``supermassive'' strings which exist in asymptotically
619: flat space only for sufficiently high enough values of the
620: vev of the Higgs field \cite{laguna}. These were considered as being ``unphysical''
621: since they should have formed at energy scales high above the GUT scale. Since recent
622: observations indicate that we live in a universe with positive cosmological constant,
623: and since we find that the existence of our solutions is not restricted to
624: values of the Higgs field's vev being large enough, these solutions might well
625: be of relevance.
626:
627: Our numerical analysis suggests that the general behaviour of the vacuum metric
628: persists in the presence of the matter fields. E.g. comparing the location of the
629: cosmological horizon $x_0$ in dependence on the coupling constants for the ``pure'' Nielsen-Olesen
630: string and that of the vacuum solution, $x^v_0$, we find that for small $\bar\Lambda$ and/or $\gamma$,
631: $x_0$ and $x^v_0$ are nearly equal. For increasing $\bar\Lambda$ and/or $\gamma$, the difference
632: between the two increases. The reason for this is that the radius of the string core
633: becomes comparable to the radius of the cosmological horizon and thus the assumption of a ``vacuum''
634: at the cosmological horizon is not valid any longer.
635:
636: Constructing the global string and the composite defect of Nielsen-Olesen and global string,
637: we find that the existence of solutions without a singularity resulting from the
638: global string itself depends crucially
639: on the product $\beta_2 q$ which (in our rescaled coordinates) is inverse proportional to
640: the radius of the global string's core. If (for fixed $q$) $\beta_2$ is too small, the core
641: of the global string extends to outside the cosmological horizon and the
642: function $f$ reaches its vev with positive concavity. For large enough values of $\beta_2$,
643: the core of the global string lies inside the horizon and
644: our numerical results seem to indicate that a global string/composite
645: defect without the normal singularity of the global string exists.
646: However, note that
647: the removal of the global string
648: singularity which exists in asymptotically flat space
649: can be achieved only by introducing a ``new'' singularity, the curvature singularity
650: at the horizon.
651: \\
652: \\
653: \\
654: \\
655: \\
656: {\bf Acknowledgement} YB acknowledges the Belgian FNRS
657: for financial support. BH was supported by an EPSRC grant.\\
658: \\
659: \\
660: {\bf Note added} After finishing the manuscript, B. Linet has brought to our intention
661: his paper ``The static, cylindrically symmetric strings in general relativity with cosmological
662: constant'' [J. Math. Phys. {\bf 27} (1986), pp. 1817-1818], in which he discusses the vacuum
663: solutions (\ref{NON}) and (\ref{NOL}) ``re-found'' by us.
664:
665:
666:
667: \begin{thebibliography}{99}
668: \bibitem{shell} A. Vilenkin and E. P. S. Shellard, {\it Cosmic strings and other topological
669: defects}, Cambridge University Press, Cambridge, England, 1994.
670: \bibitem{kibble} T. W. B. Kibble and M. Hindmarsh, Rep. Progr. Phys. {\bf 58}, 477 (1995).
671: \bibitem{no} H. B. Nielsen and P. Olesen, Nucl. Phys. {\bf B61}, 45 (1973).
672: \bibitem{Gar} D. Garfinkle, Phys. Rev. {\bf D32}, 1323 (1985).
673: \bibitem{laguna} P. Laguna and D. Garfinkle, Phys. Rev. {\bf D40}, 1011 (1989).
674: \bibitem{Christensen} M. Christensen, A. L. Larsen and Y. Verbin, Phys. Rev.
675: {\bf D60}, 125012 (1999).
676: \bibitem{yves} Y. Brihaye and M. Lubo, Phys. Rev. {\bf D62}, 085004 (2000).
677: \bibitem{cohen} A. G. Cohen and D. B. Kaplan, Phys. Lett. {\bf B215}, 67 (1988).
678: \bibitem{harari} D. Harari and P. Sikivie, Phys. Rev. {\bf D37}, 3438 (1988).
679: \bibitem{Gibbons} G. W. Gibbons, M. E. Ortiz and F. Ruiz Ruiz, Phys. Rev.
680: {\bf D39}, 1546 (1989).
681: \bibitem{gregory1} R. Gregory, Phys. Lett. {\bf B215}, 663 (1988).
682: \bibitem{vilenkin} M. Barriola and A. Vilenkin, Phys. Rev. Lett. {\bf 63}, 341 (1989);
683: D. Harari and C. Lousto, Phys. Rev. {\bf D42}, 2626 (1990).
684: \bibitem{gregory2} R. Gregory, Phys. Rev. {\bf D54}, 4955 (1996).
685: \bibitem{super} S. Perlmutter et al., Astrophys. J. {\bf 517}, 565 (1999);
686: A. G. Riess et al., Astron. J. {\bf 116}, 1009 (1998).
687: \bibitem{strominger} A. Strominger, JHEP {\bf 0110}, 034 (2001);
688: JHEP {\bf 0111}, 049 (2001).
689: \bibitem{GM} A. M. Gezelbash and R. B. Mann, Phys. Lett {\bf B537}, 329 (2002).
690: \bibitem{bbh} E. R. Bezerra de Mello, Y. Brihaye and B. Hartmann,
691: Phys. Rev. {\bf D67}, 045015 (2003).
692: \bibitem{by} Y. Brihaye and B. Hartmann, Phys. Rev. {\bf D66}, 064018 (2002).
693: \bibitem{spi} J. Spinelly, U de Freitas and E. R. Bezerra de Mello, Phys. Rev.
694: {\bf D66}, 024018 (2002).
695: \bibitem{kramer} D. Kramer, H. Stephani, E. Herlt and M. MacCallum, {\it Exact solutions of Einstein's field equations}, Cambridge University Press, Cambridge, England, 1980.
696:
697: \end{thebibliography}
698:
699:
700:
701:
702:
703:
704:
705: \newpage
706:
707: \begin{figure}
708:
709: \centering
710:
711: \epsfysize=10cm
712:
713: \mbox{\epsffile{bbh3.ps}}
714:
715: \caption{The value of the dimensionless coordinate $x$ at which a cosmological
716: horizon appears, $x=x_0$, is given for the Nielsen-Olesen string as function
717: of $\gamma$ for $n=1$ (dashed) and $n=2$ (solid) with $\alpha=1.0$ and $\bar\Lambda=0.005$. }
718:
719: \end{figure}
720:
721:
722:
723: \newpage
724:
725: \begin{figure}
726:
727: \centering
728:
729: \epsfysize=10cm
730:
731: \mbox{\epsffile{bbh4.ps}}
732:
733: \caption{The profiles of the metric functions $N(x)$, $L(x)$ and the profiles of
734: the matter field functions $P(x)$ and $h(x)$ are shown for $\alpha=1.0$,
735: $\bar\Lambda=0.005$, $n=1$ and $\gamma=1.5$ (dashed) and $\gamma=1.7$ (solid), respectively. }
736:
737: \end{figure}
738:
739:
740:
741: \newpage
742:
743: \begin{figure}
744:
745: \centering
746:
747: \epsfysize=10cm
748:
749: \mbox{\epsffile{bbh1.ps}}
750:
751: \caption{The value of the dimensionless coordinate $x$ at which a cosmological
752: horizon appears, $x=x_0$, is given for the Nielsen-Olesen string
753: as function of $\bar\Lambda$ for
754: $n=1$ and $n=2$ with $\alpha=\gamma=0.5$. For comparison, also the value
755: $x^v_0=\pi/\sqrt{3\bar\Lambda}$ for the vacuum solution is given. }
756:
757: \end{figure}
758:
759:
760:
761: \newpage
762:
763: \begin{figure}
764:
765: \centering
766:
767: \epsfysize=10cm
768:
769: \mbox{\epsffile{bbh2.ps}}
770:
771: \caption{The profiles of the metric functions $N(x)$ and $L(x)$
772: are shown for $\alpha=\gamma=0.5$,
773: $\bar\Lambda=0.01$ and vorticity $n=2$. We compare the Nielsen-Olesen solution
774: (solid) with the corresponding vacuum solution (dashed).}
775: \end{figure}
776:
777: \newpage
778: \begin{figure}
779: \centering
780: \epsfysize=10cm
781: \mbox{\epsffile{bbh5.ps}}
782: \caption{The profile of the Goldstone field function $f(x)$
783: is shown for the composite defect with $\alpha=\gamma=\beta_1^2=1$, $\beta_3^2=0$, $\bar\Lambda=0.001$, $q=0.1$,
784: vorticity $n=1$ and different values of $\beta_2^2$.}
785: \end{figure}
786:
787:
788:
789:
790:
791:
792: \end{document}
793:
794:
795: