• Nenhum resultado encontrado

Redshift Drift: Distinguishing Between ΛCDM and LTB Models

4.4 Redshift Drift: Distinguishing BetweenΛCDM and LTB Models

each∆v, with a standard deviation given by Eq. (4.10).

With these new data points it was performed a GP reconstruction, with the aim of reconstruct the theoretical function of∆v. The best-fit function found if given by

∆v=−0.182z2−0.677z, (4.11) withR2≈0.6685. These results can be found in Figure 4.11.

4.4.2 Lemaître–Tolman–Bondi models

After obtain results for theΛCDM model, we were interested in understandiing if the redshift drift for LTB models varies in a different way when comparing with theΛCDM model. For this reason, in this subsection are presented the results for both CFL and GBH models.

4.4.2.1 Clifton-Ferreira-Land model

Model Parameters

h A w(Mpc)

Fiducial 0.72 0.85 1×103 M1 0.62 0.85 1×103 M2 0.82 0.85 1×103

M3 0.72 0.8 1×103

M4 0.72 0.9 1×103

M5 0.72 0.85 0.5×103 M6 0.72 0.85 1.5×103

Table 4.4: Table with the values for the three initial parameters of the CFL models in study, both the fiducial and the six toy models. The three initial parameters are: the dimensionless Hubble constant,h, the amplitude scale of the curvature,A, and a curvature parameter,w.

The results presented here, as well as the results presented for the shear estimator, have a curvature profile (a) (see Figure 2.1). Taking this into account, we fixed a fiducial model withh=0.72,A=0.85 andw=1×103and created six additional models. Since CFL models have three initial parameters, we created the six models, with parameters up and down, i.e., around, this fiducial model parameters. We present all model parameters in Table 4.4.

In Figure 4.12, it is shown the comparison of the redshift drift for the fiducial and the six additional models considered in here. In each subplot it is represented the variation of only one parameter. We can conclude by looking at that figure that all of the CFL model parameters are sensitive to the redshift drift. Larger differences between models appear at high redshifts, so depending on the quality of future experiments, high redshift observations may be the best strategy to discriminate between models. It is also interesting to see that a bigger change between models happens when the dimensionless Hubble constant varies. To constrain bothAandwit would be necessary very precise data.

As stated before, we are interested in studying the differences betweenΛCDM and LTB models for the redshift drift. Figure 4.13 shows the redshift drift for both types of model, where we only considered the fiducial CFL model. We see an obvious difference between the two models. This is particularly striking at high redshifts (z>1) and that is why our proposed list of redshifts for targeted observations should be able to distinguish between these two classes of models.

4.4 Redshift Drift: Distinguishing BetweenΛCDM and LTB Models

(a) (b)

(c)

Figure 4.12: The redshift drift per year for the fiducial CFL and the eight toy models. The fiducial CFL model has the following parameters: h=0.72,A=0.85 andw=1×103. Each subplot represent only a change in a variable (a) the dimensionless Hubble constant,h, (b) the amplitude scale of the curvature,A, and (c) a curvature parameter,w. The black line represents the fiducial CFL model. The other colors, blue and orange, represent the six toy models as it is shown in the legend of the figure.

Figure 4.14 shows the velocity shift after a period of 10 years for theΛCDM model (black dashed line) and fiducial CFL model (green solid line). We can also see the FLRW theoretical (orange points) and mock (blue points) data. The black solid line represents the GP reconstruction taking into account the FLRW mock data.

Looking at Figure 4.14 it becomes clear that if the Universe follows the FLRW mock data generated here, it would be very easy to discard CFL models in our proposed range of redshift, since, typically, CFL models are severalσaway from the best GP function reconstruction obtained from the mock data.

4.4 Redshift Drift: Distinguishing BetweenΛCDM and LTB Models

Figure 4.13: The redshift drift per year for the fiducial CFL model and theΛCDM model. The black line represents the fiducial CFL model. The blue line represents theΛCDM model.

Figure 4.14: The velocity shift for theΛCDM model after a period of 10 years. The orange points represent the theoretical data for theΛCDM model with the error bars estimated from Eq. (4.10). The blue points represent the mock data for theΛCDM model with the error bars estimated from Eq. (4.10). The black solid line represents the GP reconstruction for theΛCDM model using mock data. The black dashed line represents the theoretical velocity shift for theΛCDM model. The green line represents the fiducial CFL model. The grey interval correspond to the confidence level of 1σ. For the confidence level of 1σ was used a fit of a polynomial with a degree two.

4.4 Redshift Drift: Distinguishing BetweenΛCDM and LTB Models

4.4.2.2 Garcia-Bellido-Haugbølle model

To see how the redshift drift changes in GBH models with different initial parameters, it was chosen a fiducial GBH model and then, by performing up and down variations for each parameter, there were created eigth additional models. The initial parameters for each of the eight models, as well as for the fiducial GBH model are shown on Table 4.5.

Model Parameters

h Ωin r0(Gpc) ∆r(Gpc)

Fiducial 0.675 0.3 1.5 0.5

M1 0.55 0.3 1.5 0.5

M2 0.8 0.3 1.5 0.5

M3 0.675 0.005 1.5 0.5

M4 0.675 0.595 1.5 0.5

M5 0.675 0.3 0.05 0.5

M6 0.675 0.3 2.95 0.5

M7 0.675 0.3 1.5 0.005

M8 0.675 0.3 1.5 0.995

Table 4.5: Table with the values for the four initial parameters of the GBH models in study, both the fiducial and the eight toy models. The four intial parameters are: the dimensionless Hubble constant,h, the underdensity at the void center,in, the size of the void,r0, and the transition width of the void profile,∆r. Bothr0and∆rare in units of Gpc.

Moreover, Figure 4.15, shows the redshift drift for all the GBH models considered in this analysis.

As can be seen, the redshift drift is sensitive to two parameters, h andΩin, this means, that through observations of the redshift drift, we may constrain the value for these parameters. Concerning M5, M6, M7 and M8, these models are all on top of the fiducial model, which means thatr0and∆rare parameters that are not sensitive to redshift drift observations.

Furthermore, taking only into account the fiducial model andΛCDM model, we were interested to see if this two models can be distinguished from one another using redshift drift observations. Figure 4.16 shows the redshift drift for these two models. It is clear that both models have a different impact on the redshift drift.

Figure 4.17 shows the velocity shift after a period of 10 years for theΛCDM model (black dashed line) and fiducial GBH model (green solid line). We can also see the FLRW theoretical (orange points) and mock (blue points) data. The black solid line represents the GP reconstruction taking into account the FLRW mock data.

Again, if we consider the redshift range of the observations we propose, see Figure 4.17 one can, for sure, distinguish between these two types of models, because our GP reconstruction confidence intervals exclude one of the models (in this case the GBH model) at manyσ.

4.4 Redshift Drift: Distinguishing BetweenΛCDM and LTB Models

(a) (b)

(c) (d)

Figure 4.15: The redshift drift per year for the fiducial GBH and the eight toy models. The fiducial GBH model has the following parameters:h=0.675,in=0.3,r0=1.5 and∆r=0.5. Each subplot represent only a change in a variable (a) the dimensionless Hubble constant,h, (b) the underdensity at the void center,in, (c) the size of the void,r0, and (d) the transition width of the void profile,∆r. The black line represents the fiducial GBH model. The other colors, blue and orange, represent the eight toy models as it is shown in the legend of the figure.

4.4 Redshift Drift: Distinguishing BetweenΛCDM and LTB Models

Figure 4.16: The redshift drift per year for the fiducial GBH model and theΛCDM model. The black line represents the fiducial GBH model. The blue line represents theΛCDM model.

Figure 4.17: The velocity shift for theΛCDM model after a period of 10 years. The orange points represent the theoretical data for theΛCDM model with the error bars estimated from Eq. (4.10). The blue points represent the mock data for theΛCDM model with the error bars estimated from Eq. (4.10). The black solid line represents the GP reconstruction for theΛCDM model using mock data. The black dashed line represents the theoretical velocity shift for theΛCDM model. The green line represents the fiducial GBH model. The grey interval correspond to the confidence level of 1σ. For the confidence level of 1σ was used a fit of a polynomial with a degree two.

Chapter 5

Conclusion

The CP is a major assumption in cosmology. Although statistical isotropy is supported be observa-tions on large scales, the CP homogeneity hypothesis is a fundamental and weakly tested assumption that is at the heart of the most popular models of modern cosmology. Therefore, it is crucial to find ways to test this assumption. In order to do that, one should consider inhomogeneous models. In particular, in this thesis, to test the CP we choose to use the LTB models. Such models are isotropic, spherically symmetric and inhomogeneous models.

However, when studying cosmological models, one must take into consideration the method used to analyse the cosmological data. Usually, in cosmology, data analysis relies on the traditional methods, known as parametric methods. However, these methods are model dependent, therefore, they fit the parameters of known cosmological functions derived from a given model. For that reason and knowing that different classes of models have different functions and different parameters, it is easy to conclude that parameter constraints become model dependent quantities and, therefore, it may be very hard to understand if real progress is being made at understanding the underlying physical process behind the quantities being measured. Thus, how can we know which is the real model that best describes the Universe?

These and other questions took us to study alternative methods to analyse cosmological data. As the computational power increases, so do non-parametric methods. These alternative methods have the advantage of being model independent, as they rely only on data. This means that we do not need to define a modela priori, instead the data will tell us which is the model behind such results.

In this thesis, in order to implement non-parametric methods, we used an AI algorithm known as GP.

AI has been a powerful tool when it comes to analyse cosmological data. This makes the results more reliable since they are not biased by a model. This is essential, especially because present data reveals several tensions that may be a consequence of imposing pre-defined functions, instead of letting the data to provide us with its preferred functions.

The first thing we were concerned with was if non-parametric methods can be as precise as parametric methods. To study such topic we used the SCP Union2.1 SN Ia compilation data of the distance modulus.

Using the parametric approach we were able to find results for theΛCDM parameters with a coefficient of determination ofR2≈0.9930. On the other hand, by using the non-parametric approach we found a function for the distance modulus with a coefficient of determination of R2≈0.9879. Both methods offered high coefficients of determination showing for this exercise that non-parametric methods are comparable with parametric methods and can be an alternative to the traditional methods.

So if we want to use non-parametric methods as an alternative to analyse data, we first have to understand them. Hence, we tried to replicate the results obtained by Arjona and Nesseris in Ref. [41].

Even using the same data, we were not able to exactly replicate their results. However, this is not a surprise given the stochastic nature of GP and since both of us used different GP implementations we naturally obtained different formulas, but they are very close and within the function margins of each other. Note that this does not happen only if one uses different GP implementations, this can also happen when one performs different runs using the same implementation.

Next, we were faced with another important questions. Although non-parametric methods are able to replicate results with the same precision of parametric methods, what would be the limit noise level of data that would guarantee a good reconstruction? Are we already in the presence of such precise data? To answer these questions we reconstruct two cosmological functions: the luminosity distance and the shear estimator. For both cosmological functions we conclude that the quality of the GP reconstruction depends strongly on the noise level of the data. Only low fractions of present noise levels allow the reconstruction of the true theoretical function for the used quantities. More specifically, for the luminosity distance we were able to obtain the theoretical function until we reach noise level of 0.3σdA(z). Regarding the shear estimator, we were able to obtain its function until a noise level of 0.15σH. Thus, we must wait for a more precise generation of cosmological data to use GP as a powerful reconstruction method for these functions.

However, one should note that although the paper of Arjona and Nesseris [41] is from 2020, the data they use is much older, therefore, until this day, the precision of data has increased a lot. For this reason, one future step of this thesis, would be to take into account more precise data that already exists for this analysis. In addition, we would like to extend this analysis to other cosmological functions and for the uncoming galaxy surveys, such as the Euclid survey.

Still regarding the shear estimator, one of our future steps would be to rewrite this null test with observables more easy to measure so we could be able to perform such test.

Moreover, after a deep study of GP, we focused on another important part of this thesis. We were interested in distinguishing between FLRW and LTB models, so, we investigated the potential of using observations of the redshift drift for this purpose. We created data for what would be expected for the ΛCDM model if we could do such observation. Then, we compared with both CFL and GBH models and conclude that the LTB models used would be discarded easily from the FLRW mock data if we perform this type of observations. The redshift drift is a remarkable way to distinguish cosmological models and the power of such observation would be an amazing step to test the CP.

Since our LTB models can be easily discarded if we observe the redshift drift as it is proposed in Chapter 4, one future concern we have is to extend the range of LTB models to understand if all of them behave so distinctively from FLRW models. If so, the study of such models can be abandoned by performing this simple observation.

Bibliography

[1] Francesco Sylos Labini and Yuri V. Baryshev. “Testing the Copernican and Cosmological Prin-ciples in the local universe with galaxy surveys”. In: 2010.6, 021 (June 2010), p. 021. DOI:10.

1088/1475-7516/2010/06/021. arXiv:1006.0801 [astro-ph.CO].

[2] Håvard Alnes, Morad Amarzguioui, and Øyvind Grøn. “Inhomogeneous alternative to dark en-ergy?” In: 73.8, 083519 (Apr. 2006), p. 083519. DOI:10.1103/PhysRevD.73.083519. arXiv:

astro-ph/0512006 [astro-ph].

[3] G. Lemaître. “The Expanding Universe”. In:Gen. Rel. Grav.29 (1997), pp. 641–680.

[4] Richard C. Tolman. “Effect of Inhomogeneity on Cosmological Models”. In:Proceedings of the National Academy of Science20.3 (Mar. 1934), pp. 169–176.DOI:10.1073/pnas.20.3.169.

[5] H. Bondi. “Spherically symmetrical models in general relativity”. In: 107 (Jan. 1947), p. 410.DOI: 10.1093/mnras/107.5-6.410.

[6] Philip Bull, Timothy Clifton, and Pedro G. Ferreira. “Kinematic Sunyaev-Zel’dovich effect as a test of general radial inhomogeneity in Lemaıtre-Tolman-Bondi cosmology”. In: 85.2, 024002 (Jan. 2012), p. 024002. DOI: 10 . 1103 / PhysRevD . 85 . 024002. arXiv: 1108 . 2222 [astro-ph.CO].

[7] David Camarena et al. “The Copernican principle in light of the latest cosmological data”. In:

509.1 (Jan. 2022), pp. 1291–1302. DOI: 10 . 1093 / mnras / stab3077. arXiv: 2107 . 02296 [astro-ph.CO].

[8] David Camarena et al. “A void in the Hubble tension? The end of the line for the Hubble bubble”.

In:Classical and Quantum Gravity39.18, 184001 (Sept. 2022), p. 184001.DOI: 10.1088/1361-6382/ac8635. arXiv:2205.05422 [astro-ph.CO].

[9] Marco San Martın and Carlos Rubio. “Hubble tension and matter inhomogeneities: a theoreti-cal perspective”. In:arXiv e-prints, arXiv:2107.14377 (July 2021), arXiv:2107.14377. DOI:10.

48550/arXiv.2107.14377. arXiv:2107.14377 [astro-ph.CO].

[10] Sveva Castello, Marcus Högås, and Edvard Mörtsell. “A cosmological underdensity does not solve the Hubble tension”. In: 2022.7, 003 (July 2022), p. 003.DOI:10.1088/1475-7516/2022/07/

003. arXiv:2110.04226 [astro-ph.CO].

[11] Rong-Gen Cai et al. “Do the observational data favor a local void?” In: 103.12, 123539 (June 2021), p. 123539.DOI:10.1103/PhysRevD.103.123539. arXiv:2012.08292 [astro-ph.CO].

[12] W. D’Arcy Kenworthy, Dan Scolnic, and Adam Riess. “The Local Perspective on the Hubble Tension: Local Structure Does Not Impact Measurement of the Hubble Constant”. In: 875.2, 145 (Apr. 2019), p. 145.DOI:10.3847/1538-4357/ab0ebf. arXiv:1901.08681 [astro-ph.CO].

BIBLIOGRAPHY

[13] Benjamin L. Hoscheit and Amy J. Barger. “The KBC Void: Consistency with Supernovae Type Ia and the Kinematic SZ Effect in aΛLTB Model”. In: 854.1, 46 (Feb. 2018), p. 46.DOI:10.3847/

1538-4357/aaa59b. arXiv:1801.01890 [astro-ph.CO].

[14] Euclid Collaboration et al. “Euclid preparation. I. The Euclid Wide Survey”. In: 662, A112 (June 2022), A112.DOI:10.1051/0004-6361/202141938. arXiv:2108.01201 [astro-ph.CO].

[15] DESI Collaboration et al. “The DESI Experiment Part I: Science,Targeting, and Survey De-sign”. In:arXiv e-prints, arXiv:1611.00036 (Oct. 2016), arXiv:1611.00036. arXiv:1611.00036 [astro-ph.IM].

[16] K. C. Chambers et al. “The Pan-STARRS1 Surveys”. In:arXiv e-prints, arXiv:1612.05560 (Dec.

2016), arXiv:1612.05560. arXiv:1612.05560 [astro-ph.IM].

[17] Scott Dodelson.Modern Cosmology. Academic Press, Elsevier Science, 2003.

[18] Austin Joyce et al. “Beyond the cosmological standard model”. In: 568 (Mar. 2015), pp. 1–98.

DOI:10.1016/j.physrep.2014.12.002. arXiv:1407.0059 [astro-ph.CO].

[19] Wendy L. Freedman. “Measurements of the Hubble Constant: Tensions in Perspective”. In: 919.1, 16 (Sept. 2021), p. 16.DOI:10.3847/1538-4357/ac0e95. arXiv:2106.15656 [astro-ph.CO].

[20] Gaetano Lambiase et al. “Testing dark energy models in the light ofσ _8 tension”. In:European Physical Journal C79.2, 141 (Feb. 2019), p. 141.DOI:10.1140/epjc/s10052- 019- 6634- 6.

arXiv:1804.07154 [astro-ph.CO].

[21] Juan Garcia-Bellido and Troels Haugbølle. “Confronting Lemaitre Tolman Bondi models with observational cosmology”. In: 2008.4, 003 (Apr. 2008), p. 003. DOI:10 . 1088 / 1475 - 7516 / 2008/04/003. arXiv:0802.1523 [astro-ph].

[22] Pantelis S. Apostolopoulos et al. “Cosmological acceleration and gravitational collapse”. In:

2006.6, 009 (June 2006), p. 009. DOI:10 . 1088 / 1475 - 7516 / 2006 / 06 / 009. arXiv: astro -ph/0603234 [astro-ph].

[23] Juan Garcıa-Bellido and Troels Haugbølle. “The radial BAO scale and cosmic shear, a new observ-able for inhomogeneous cosmologies”. In: 2009.9, 028 (Sept. 2009), p. 028.DOI: 10.1088/1475-7516/2009/09/028. arXiv:0810.4939 [astro-ph].

[24] Timothy Clifton, Pedro G. Ferreira, and Kate Land. “Living in a Void: Testing the Copernican Principle with Distant Supernovae”. In: 101.13, 131302 (Sept. 2008), p. 131302.DOI:10.1103/

PhysRevLett.101.131302. arXiv:0807.1443 [astro-ph].

[25] Allan Sandage. “The Change of Redshift and Apparent Luminosity of Galaxies due to the Decel-eration of Selected Expanding Universes.” In: 136 (Sept. 1962), p. 319.DOI:10.1086/147385.

[26] Abraham Loeb. “Direct Measurement of Cosmological Parameters from the Cosmic Deceleration of Extragalactic Objects”. In: 499.2 (June 1998), pp. L111–L114.DOI:10.1086/311375. arXiv:

astro-ph/9802122 [astro-ph].

[27] Claudia Quercellini et al. “Real-time cosmology”. In: 521.3 (Dec. 2012), pp. 95–134.DOI:10.

1016/j.physrep.2012.09.002. arXiv:1011.2646 [astro-ph.CO].

[28] R. Codur and C. Marinoni. “Redshift drift in radially inhomogeneous Lemaıtre-Tolman-Bondi spacetimes”. In:arXiv e-prints, arXiv:2107.04868 (July 2021), arXiv:2107.04868. arXiv:2107.

04868 [gr-qc].

BIBLIOGRAPHY

[29] Sanjib Sharma. “Markov Chain Monte Carlo Methods for Bayesian Data Analysis in Astronomy”.

In: 55.1 (Aug. 2017), pp. 213–259.DOI:10.1146/annurev- astro- 082214- 122339. arXiv:

1706.01629 [astro-ph.IM].

[30] Jonathan Goodman and Jonathan Weare. “Ensemble samplers with affine invariance”. In: Com-munications in Applied Mathematics and Computational Science5.1 (Jan. 2010), pp. 65–80.DOI: 10.2140/camcos.2010.5.65.

[31] Kenneth Levenberg. “A Method for the Solution of Certain Non-Linear Problems in Least Squares”. In:The Quarterly of Applied Mathematics2 (1944), pp. 164–168.

[32] Donald W. Marquardt. “An Algorithm for Least-Squares Estimation of Nonlinear Parameters”.

In:Journal of the Society for Industrial and Applied Mathematics11.2 (1963), pp. 431–441.DOI: 10.1137/0111030. eprint:https://doi.org/10.1137/0111030.URL:https://doi.org/

10.1137/0111030.

[33] Planck Collaboration et al. “Planck 2018 results. VI. Cosmological parameters”. In: 641, A6 (Sept.

2020), A6.DOI:10.1051/0004-6361/201833910. arXiv:1807.06209 [astro-ph.CO].

[34] G. James et al.An Introduction to Statistical Learning: with Applications in R. Springer Texts in Statistics. Springer New York, 2014.ISBN: 9781461471370.URL:https://books.google.pt/

books?id=at1bmAEACAAJ.

[35] Shu-heng Chen. “Genetic Algorithms and Genetic Programming in Computational Finance.” In:

J. Artificial Societies and Social Simulation7 (Jan. 2004).

[36] Amir Dehsarvi and Stephen L. Smith. “Classification of Resting-State fMRI using Evolutionary Algorithms: Towards a Brain Imaging Biomarker for Parkinson’s Disease”. In: arXiv e-prints, arXiv:1910.05378 (Oct. 2019), arXiv:1910.05378. arXiv:1910.05378 [cs.NE].

[37] Christian Loftis et al. “Lattice Thermal Conductivity Prediction Using Symbolic Regression and Machine Learning”. In:Journal of Physical Chemistry A 125.1 (Jan. 2021), pp. 435–450. DOI: 10.1021/acs.jpca.0c08103. arXiv:2008.03670 [cond-mat.mtrl-sci].

[38] Michelle Ntampaka et al. “The Role of Machine Learning in the Next Decade of Cosmology”. In:

51.3, 14 (May 2019), p. 14. arXiv:1902.10159 [astro-ph.IM].

[39] John R. Koza. Genetic Programming: On the Programming of Computers by Means of Natural Selection. Cambridge, MA, USA: MIT Press, 1992.ISBN: 0-262-11170-5.

[40] Savvas Nesseris and Juan Garcıa-Bellido. “A new perspective on dark energy modeling via genetic algorithms”. In: 2012.11, 033 (Nov. 2012), p. 033. DOI:10.1088/1475- 7516/2012/11/033.

arXiv:1205.0364 [astro-ph.CO].

[41] Rubén Arjona and Savvas Nesseris. “What can machine learning tell us about the background expansion of the Universe?” In: Phys. Rev. D101 (12 June 2020), p. 123525. DOI:10 . 1103 / PhysRevD . 101 . 123525. URL:https : / / link . aps . org / doi / 10 . 1103 / PhysRevD . 101 . 123525.

[42] N. Suzuki et al. “The Hubble Space Telescope Cluster Supernova Survey. V. Improving the Dark-energy Constraints above z > 1 and Building an Early-type-hosted Supernova Sample”.

In: 746.1, 85 (Feb. 2012), p. 85. DOI:10 . 1088 / 0004 - 637X / 746 / 1 / 85. arXiv: 1105 . 3470 [astro-ph.CO].

BIBLIOGRAPHY

[43] Varun Sahni, Arman Shafieloo, and Alexei A. Starobinsky. “Two new diagnostics of dark energy”.

In: 78.10, 103502 (Nov. 2008), p. 103502.DOI:10.1103/PhysRevD.78.103502. arXiv:0807.

3548 [astro-ph].

[44] S. Nesseris et al. “Euclid: Forecast constraints on consistency tests of the ΛCDM model”. In:

660, A67 (Apr. 2022), A67. DOI: 10 . 1051 / 0004 - 6361 / 202142503. arXiv: 2110 . 11421 [astro-ph.CO].

[45] Luca Pasquini et al. “CODEX: Measuring the Expansion of the Universe (and beyond)”. In:The Messenger122 (Dec. 2005), pp. 10–14.

[46] Luca Pasquini et al. “CODEX: measuring the acceleration of the universe and beyond”. In:The Scientific Requirements for Extremely Large Telescopes. Ed. by P. Whitelock, M. Dennefeld, and B. Leibundgut. Vol. 232. Jan. 2006, pp. 193–197.DOI:10.1017/S1743921306000482.

[47] Ruth Lazkoz, Iker Leanizbarrutia, and Vincenzo Salzano. “Forecast and analysis of the cosmolog-ical redshift drift”. In:European Physical Journal C78.1, 11 (Jan. 2018), p. 11.DOI:10.1140/

epjc/s10052-017-5479-0. arXiv:1712.07555 [astro-ph.CO].

Appendices

Appendix A

Hubble Parameter Code

In this Appendix, there is our code for the Hubble Parameter, which was used in Chapter 4 (Sec.

4.2.1).

1 #!/ usr / bin / env p y t h o n 3 2 # * c o d i n g : utf 8 * -3 """

4 C r e a t e d on Fri Nov 26 1 5 : 0 8 : 0 0 2 0 2 1 5

6 @ a u t h o r : M a r i a G o n c a l v e s 7 """

8

9 f r o m g p l e a r n . g e n e t i c i m p o r t S y m b o l i c R e g r e s s o r 10 i m p o r t m a t p l o t l i b . p y p l o t as plt

11 i m p o r t n u m p y as np 12 i m p o r t g r a p h v i z 13 f r o m s y m p y i m p o r t * 14 i m p o r t m at h

15 16 17

18 # R e a l d a t a 19 h _ d a t a = []

20 z _ d a t a =[]

21 s i g m a =[]

22 23

24 f = o p e n ( ’ z - h ( z ) ’ , ’r ’) 25 i =0

26 for l i n e in f :

27 if i <= 35:

28 s = l i n e . s p l i t ()

29 h _ d a t a . a p p e n d ( f l o a t ( s [ 1 ] ) ) 30 z _ d a t a . a p p e n d ( f l o a t ( s [ 0 ] ) ) 31 s i g m a . a p p e n d ( f l o a t ( s [ 2 ] ) )

32 i +=1

33 e l s e:

34 b r e a k

35

36 f . c l o s e () 37

38 39

40 z = np. r e s h a p e ( z_data , (36 ,1)) 41

42 43 44 45

46 # S y m b o l i c R e g r e s s o r C l a s s

47 gp = S y m b o l i c R e g r e s s o r ( p o p u l a t i o n _ s i z e =100 , g e n e r a t i o n s =1000 ,

48 t o u r n a m e n t _ s i z e =10 , s t o p p i n g _ c r i t e r i a = 0 . 0 0 0 1 ,

49 c o n s t _ r a n g e =( -1.0 , 1.0) , i n i t _ d e p t h =(2 , 10) ,

50 i n i t _ m e t h o d = ’ h a l f and half ’ ,

51 f u n c t i o n _ s e t =( ’ add ’ , ’ sub ’ , ’ mul ’ , ’ div ’) ,

52 m e t r i c = ’ rmse ’ , p a r s i m o n y _ c o e f f i c i e n t =0.01 ,

53 p _ c r o s s o v e r =0.5 , p _ s u b t r e e _ m u t a t i o n =0.01 ,

54 p _ h o i s t _ m u t a t i o n =0.01 , p _ p o i n t _ m u t a t i o n =0.01 ,

55 p _ p o i n t _ r e p l a c e =0.01 , m a x _ s a m p l e s =1 ,

56 f e a t u r e _ n a m e s = ’ z ’ , w a r m _ s t a r t =False,

57 l o w _ m e m o r y =False, n _ j o b s =1 ,

58 v e r b o s e =1 , r a n d o m _ s t a t e = 2 3 )

59

60 # Fit the G e n e t i c P r o g r a m a c c o r d i n g to z and h _ d a t a 61 gp . fit ( z , h _ d a t a )

62 63 64 65 66 67

68 # o p e r a t i o n s c o n v e r t i o n for p r i n t i n g f i n a l f u n c t i o n 69 c o n v e r t e r ={ ’ add ’: l a m b d a x , y : x + y ,

70 ’ sub ’: l a m b d a x , y : x - y , 71 ’ mul ’: l a m b d a x , y : x * y , 72 ’ div ’: l a m b d a x , y : x / y } 73

74

75 # f i n a l f u n c t i o n s i m p l i f i e d

76 f u n c t i o n = s y m p i f y ( str ( gp . _ p r o g r a m ) , l o c a l s = c o n v e r t e r ) 77 p r i n t( f u n c t i o n )

78 79

80 # h p r e d i c t e d by the a l g o r i t h m t a k i n g i n t o a c c o u n t z 81 h = gp . p r e d i c t ( z )

82

83 # c o e f f i c i e n t of d e t e r m i n a t i o n 84 p r i n t( gp . s c o r e ( z , h _ d a t a )) 85

86 87 88 89

90 # 1 s i g m a i n t e r v a l c a l c u l a t i o n 91 dfp =[]

92 dfn =[]

93 for i in r a n g e ( len ( z )):

94 eps = 0 . 0 0 0 1 95 d e l t a _ f =1

96 c = m a t h . erf ( 1 / ( m a t h . s q r t ( 2 ) ) ) 97

98 f1 = ( 1 / 2 ) * ( ma t h . erf (( d e l t a _ f + h [ i ] - h _ d a t a [ i ])/

99 ( m a t h . s q r t ( 2 ) * s i g m a [ i ]))

100 + m a t h . erf (( delta_f - h [ i ]+ h _ d a t a [ i ])/

101 ( m a t h . s q r t ( 2 ) * s i g m a [ i ]) ) )

102 103

104 w h i l e abs ( f1 - c ) > eps : 105

106

107 if f1 > c :

108 delta_f - = 0 . 0 0 1

109

110 if f1 < c :

111 d e l t a _ f + = 0 . 0 0 1

112

113 f1 = ( 1 / 2 ) * ( m a t h . erf (( d e l t a _ f + h [ i ] - h _ d a t a [ i ])/

114 ( m a t h . s q r t ( 2 ) * s i g m a [ i ]))

115 + m a t h . erf (( delta_f - h [ i ]+ h _ d a t a [ i ])/

116 ( m a t h . s q r t ( 2 ) * s i g m a [ i ]) ) )

117 118 119

120 dfp . a p p e n d ( d e l t a _ f + h [ i ]) 121 dfn . a p p e n d ( h [ i ] - d e l t a _ f ) 122

123 124 125 126

127 ## p o l i n o m i o CI 128

129 zz = z . r a v e l () 130

131 # c o e f f i c i e n t v a l u e s

132 p o l y _ d f p = np. p o l y f i t ( zz , dfp , deg =2) 133 p o l y _ d f n = np. p o l y f i t ( zz , dfn , deg =2) 134

135

136 # p o l y n o m i a l p o i n t s

137 p _ d f p =np. p o l y v a l ( p o l y _ d f p , zz ) 138 p _ d f n =np. p o l y v a l ( p o l y _ d f n , zz ) 139

140 ###

141 142 143 144 145 146 147

148 h _ l c d m =[]

149 for i in r a n g e ( len ( z )):

150

151 h _ l c d m . a p p e n d ( 6 9 . 2 7 * m a t h . s q r t ( 0 . 6 8 4 7 + 0 . 3 1 5 3 * ( 1 + z [ i ] ) * * 3 ) ) 152

153 154

155 h _ p a p e r =[]

Documentos relacionados