Translate

Monday, November 30, 2015

Basically, matter is the rarest thing in the universe (or at least it used to be)

We are bounded in a nutshell of Infinite space: Blog Post #35, Worksheet # 11.1, Problem #3: Basically, matter is the rarest thing in the universe (or at least it used to be)

3. Baryon-to-photon ratio of our universe.
(a) Despite the fact that the CMB has a very low temperature (that you have calculated above), the number of photons is enormous. Let us estimate what that number is. Each photon has energy \(h\nu\). From the Plank Spectrum Equation, figure out the number density, \(n_\nu\), of the photon per frequency interval \(d\nu\). Integrate over \(d\nu\)to get an expression for total number density of photon given temperature T. Now you need to keep all factors, and use the fact that \[\int_0^\infty \frac{x^2}{e^x - 1} dx \approx 2.4.\]

(b) Use the following values for the constants: \(k_B = 1.38 \times 10^{-16} erg~ K^{-1}, c = 3.00 \times 10^{10} cm ~s^{-1}, h = 6.62 \times 10^{-27} erg~s\) , and use the temperature of CMB today that you have computed from 2d), to calculate the number density of photon today in our universe today (i.e. how many photons per cubic centimeter?)

(c) Let us calculate the average baryon number density today. In general, baryons refer to protons or neutrons. The present-day density (matter + radiation + dark energy) of our Universe is \(9.2 \times 10^{-30} g~cm^{-3}\) . The baryon density is about 4% of it. The masses of proton and neutron are very similar \((\approx 1.7 \times  10^{-24} G )\).
What is the number density of baryons?

(d) Divide the above two numbers, you get the baryon-to-photon ratio. As you can see, our universe contains much more photons than baryons (proton and neutron).

(a) Knowing the energy of a single photon, a simple scale value defines all photon’s energy: \[E = h\nu \] \[E_n = n\cdot h\nu \] \[n_\nu = \frac{E_{n_\nu}}{h\nu},\] which can be defined with the equations for energy density as: \[ n_\nu = \frac{u_\nu}{h\nu},\] so \[u_\nu d\nu = \frac{8\pi h_P \nu^3}{c^3} \frac{1}{e^{\frac{h_P \nu}{k_B T}} - 1} d\nu,\]becomes: \[\frac{u_\nu}{h\nu} d\nu = \frac{1}{h\nu} \cdot \frac{8\pi h_P \nu^3}{c^3} \frac{1}{e^{\frac{h_P \nu}{k_B T}} - 1} d\nu,\] \[n_\nu d\nu =  \frac{8\pi \nu^2}{c^3} \frac{1}{e^{\frac{h_P \nu}{k_B T}} - 1} d\nu,\] and integrating it for all frequencies: \[\int_0 ^\infty n_\nu d\nu = \int_0 ^\infty \frac{8\pi \nu^2}{c^3} \frac{1}{e^{\frac{h_P \nu}{k_B T}} - 1} d\nu,\] \[n_{all \nu} =  \frac{8\pi}{c^3} \left(\frac{k_B T}{h_P}\right)^3 \int_0^\infty \frac{\nu^2}{e^{\frac{h_P \nu}{k_B T}} - 1} d\nu,\] and knowing a particular equality established by the problem: \[\int_0^\infty \frac{x^2}{e^x - 1} dx \approx 2.4 ,\] we have the final description of the number of photons as: \[n_{all ~\nu} =  \frac{8\pi}{c^3} \left(\frac{k_B T}{h_P}\right)^3 (2.4)\]

(b) And now plugging in the values given in the problem, we have: \[n_{all~ \nu} =  \frac{8\pi}{c^3} \left(\frac{k_B T}{h_P}\right)^3 (2.4)\] \[k_B = 1.38 \times 10^{-16} erg~ K^{-1}, \] \[c = 3.00 \times 10^{10} cm ~s^{-1},\] \[h = 6.62 \times 10^{-27} erg~s, \] so:  
\[n_{all~ \nu} =  \frac{8\pi}{3.00 \times (10^{10} cm ~s^{-1})^3} \left(\frac{1.38 \times 10^{-16} erg~ K^{-1} \cdot 2.7K}{ 6.62 \times 10^{-27} erg~s }\right)^3 (2.4),\] and thus the density of photons is: \[n_{all ~\nu} = 4 \times 10^2 ~photons/cm^3\]

(c)  Being given the density of mass in the universe, we can now solve for baryon density in particular: \[\rho_{universe} = 9.2 \times 10^{-30} g~cm^{-3}\] \[\rho_{baryons} = 0.04 \times \rho_{universe} ,\] and knowing the mass of a baryon, we can solve for the number density of baryons: \[\rho_{baryons} = 3.68*10^{-31} g~cm^{-3}\] \[n_{baryons} = \rho_{baryons} / m_{baryon}\] \[ n_{baryons} =3.68*10^{-31} g~cm^{-3} / 1.7 * 10^{-24} \frac{g}{baryon}\] \[ n_{baryon} = 2.16*10^{-7} \frac{baryons}{cm^3}\]


(d) Simply doing what the problem establishes, we have: \[n_{all \nu} = 4 \times 10^2 ~photons/cm^3\] \[ n_{baryon} = 2.16471*10^{-7} \frac{baryons}{cm^3}\] \[n_{all ~\nu}/ n_{baryon} = 1.85*10^9 \frac{photons}{baryons},\] and indeed, there are billions of photons for every baryon. 

Energy that was there 13.68 billion years ago

We are bounded in a nutshell of Infinite space: Blog Post #34, Worksheet # 11.1, Problem #2: Energy that was there 13.68 billion years ago

2. Cosmic microwave background. One of the successful predictions of the Big Bang model is the cosmic microwave background (CMB) existing today. In this exercise let us figure out the spectrum and temperature of the CMB today.
In the Big Bang model, the universe started with a hot radiation-dominated soup in thermal equilibrium. In particular, the spectrum of the electromagnetic radiation (the particle content of the electromagnetic radiation is called photon) satisfies The Plank Spectrum. At about the redshift \(z \approx 1100\) when the universe had the temperature \(T \approx 3000K\), almost all the electrons and protons in our universe are combined and the universe becomes electromagnetically neutral. So the electromagnetic waves (photons) no longer get absorbed or scattered by the rest of the contents of the universe. They started to propagate freely in the universe until reaching our detectors. Interestingly, even though the photons are no longer in equilibrium with its environment, as we will see, the spectrum still maintains an identical form to the Planck spectrum, albeit characterized by a different temperature.

(a) If the photons was emitted at redshift z with frequency \(\nu\), what is its frequency  \(\nu^{\prime}\) today?

(b) If a photon at redshift z had the energy density \(u_\nu d\nu\), what is its energy density \(u_{\nu^\prime}d\nu^\prime\) today? (Hint, consider two effects: 1) the number density of photons is diluted; 2) each photon is redshifted so its energy, \(E = hv\), is also redshifted.)

(c) Plug in the relation between \(\nu\) and \(\nu^\prime\) into the Planck spectrum: \[u_\nu d\nu = \frac{8\pi h_P \nu^3}{c^3} \frac{1}{e^{\frac{h_P \nu}{k_B T}} - 1} d\nu,\] and also multiply it with the overall energy density dilution factor that you have just figured out above to get the energy density today. Write the final expression as the form \(u_{\nu^\prime}d\nu^\prime\). What is \(u_{\nu^\prime}\)? This is the spectrum we observe today. Show that it is exactly the same Planck spectrum, except that the temperature is now \(T^\prime = T(1+z)^{-1}\)  .

(d) As you have just derived, according to Big Bang model, we should observe a black body radiation with temperature \(T^\prime \) filled in the entire universe. This is the CMB. Using the information given at the beginning of this problem, what is this temperature \(T^\prime\) today? (This was indeed observed first in 1964 by American radio astronomers Arno Penzias and Robert Wilson, who were awarded the 1978 Nobel Prize.)

(a) Starting off with the normal redshift equation, it can be turned into a redshift equation that describes the same value with photon frequencies: \[z = \frac{\lambda_{observed} - \lambda_{emitted}}{\lambda_{emitted}} \] \[\lambda = \frac{c}{\nu}\] \[z = \frac{\frac{c}{\nu_{observed}} - \frac{c}{\nu_{emitted}}}{\frac{c}{\nu_{emitted}}} ,\] and simplyfing the equation a bit to find a sole value for the \(\nu^\prime\): \[z = \frac{\nu_{emitted}} {\nu_{observed}} - 1 \] \[z + 1= \frac{\nu_{emitted}} {\nu^\prime} \]\[ \nu^\prime= \frac{\nu_{emitted}}{ z + 1},\] so we now see how the frequency seen today is always larger than the original frequency with which the photon was emitted.

(b)If we know a basic tenant of cosmology, the way the scale factor changes with redshift: \[a =\frac{1}{1+z},\] we can use it to describe the rest of the proportions which describe how the universe changes over periods of time. Knowing that \[ \lambda \propto a(t) \] from \[z = \frac{\lambda_{observed} - \lambda_{emitted}}{\lambda_{emitted}} \] \[z + 1= \frac{\lambda_{observed}} {\lambda_{emitted}} \] and focusing in on \(\lambda_{observed}\), we know how  \[ \lambda \propto a(t) \] and how\(\lambda = \frac{c}{\nu}\) so the relationship can be described as: \[ \frac{1}{\nu} \propto a(t) \] \[ \nu \propto \frac{1}{a(t)} \] which can be rewritten as:\[ \nu^\prime \propto \frac{1}{a^\prime (t)} .\] Another aspect that the problem describes is how volume is related to the scale factor, which is simply: \[V \sim a(t)^3,\] and so \[V^\prime \sim a(t)^{\prime~ 3}.\] With these, we can start doing some squiggle math to correctly describe the relationships of the variables: \[E\sim \nu\] \[E^\prime\sim \nu^prime,\] and the energy density of the universe is described as \[\rho_E = \frac{E}{V},\] so it can also be: \[\rho_E \sim \frac{\nu}{V}\] \[\rho_E \sim \frac{\frac{1}{a}}{\frac{1}{a^3}}\] \[\rho_E = \frac{1}{a^4},\] which is just scaled as anything else has for this problem: \[\rho_E^\prime = \frac{1}{a^{\prime 4}},\] and since the scale factor in the present is one: \[\rho_E^\prime = 1 ,\] and thus:  \[\frac{\rho_E}{\rho_E^\prime} \sim \frac{\frac{1}{a^4}}{1}\] \[\frac{\rho_E}{\rho_E^\prime} \sim \frac{1}{a^4},\] and applying the definition of the scale factor from redshift: \[\frac{\rho_E}{\rho_E^\prime} \sim (1+z)^4,\] so the energy density is what the Plank Spectrum calculates: \[\frac{u_\nu d\nu}{u_{\nu^\prime} d\nu^\prime} \sim (1+z)^4\] \[u_\nu d\nu \sim u_{\nu^\prime} d\nu^\prime (1+z)^4 \] \[ u_{\nu^\prime} d\nu^\prime \sim \frac{u_\nu d\nu}{(1+z)^4},\] which is how energy density scales with redshift (i.e. time).

(c) Knowing the Plank Spectrum equation: \[u_\nu d\nu = \frac{8\pi h_P \nu^3}{c^3} \frac{1}{e^{\frac{h_P \nu}{k_B T}} - 1} d\nu,\] and knowing how it scales with time (redshift): \[z + 1= \frac{\nu_{emitted}} {\nu^\prime} \] \[ \nu_{emitted} =\nu^\prime (z + 1)\]  \[ d\nu_{emitted} =d\nu^\prime (z + 1),\] we can solve for how it changes:\[u_\nu d\nu = \frac{8\pi h_P (\nu^\prime (z + 1))^3}{c^3} \frac{1}{e^{\frac{h_P (\nu^\prime (z + 1))}{k_B T}} - 1} (d\nu^\prime (z + 1))\] \[ u_{\nu^\prime} d\nu^\prime \sim \frac{u_\nu d\nu}{(1+z)^4}\] \[\frac{u_\nu d\nu}{(1+z)^4} = \frac{\frac{8\pi h_P (\nu^\prime (z + 1))^3}{c^3} \frac{1}{e^{\frac{h_P (\nu^\prime (z + 1))}{k_B T}} - 1} (d\nu^\prime (z + 1))}{ (1+z)^4},\] \[ u_{\nu^\prime} d\nu^\prime = \frac{8\pi h_P \nu^\prime3}{(z+1) c^3} \frac{1}{e^{\frac{h_P (\nu^\prime (z + 1))}{k_B T}} - 1} (d\nu^\prime (z + 1))\]  \[u_{\nu^\prime} d\nu^\prime = \frac{8\pi h_P \nu^\prime3}{c^3} \frac{1}{e^{\frac{h_P (\nu^\prime (z + 1)}{k_B T}} - 1} (d\nu^\prime ),\] and after all that algebra, we can solve for the final (z +1) by incorporating the change in temperature equation: \[T^\prime = T(1+z)^{-1},\] so the final equation becomes:  \[u_{\nu^\prime} d\nu^\prime = \frac{8\pi h_P \nu^\prime3}{c^3} \frac{1}{e^{\frac{h_P (\nu^\prime}{k_B T^\prime}} - 1} (d\nu^\prime)\]

(d) For the last part, we simply have to incorporate the definitions given in the beginning of the problem and the temperature change equation: \[T^\prime = T(1+z)^{-1}\] \[T = 3000 K\] \[ z= 1100\] \[ T^\prime = 3000K \cdot (1+1100)^{-1}\] \[T^\prime = 2.7 K,\] which is the temperature of the Cosmic Microwave background in the present. 

Monday, November 23, 2015

Relativity, the math, yeah, really, learn it, cry over it, love it, not necessarily in that order

We are bounded in a nutshell of Infinite space: Blog Post #33, Free Form #5: Relativity, the math, yeah, really, learn it, cry over it, love it, not necessarily in that order.

Following up the earlier blog post in September (http://ay17-rcordova.blogspot.com/2015/09/relativity-yes-we-are-doing-this-it-is.html), relativity is an essential part to how we’ve come to understand how the universe works, especially when considering the fact that the largest objects and forces in the universe move at speeds comparable, if not at, the speed of light. So, we shall begin with the fundamentals of how Special Relativity, as told to you by one who learned it not a month ago.
In Special Relativity, there are two basic assumptions which make up the basis of all the equations:
1.       There are no preferred inertial frames
2.       The speed of light is the same in every inertial frame
From these two basic tenets, there are three Fundamental effects that can be observed, and have been consistently tested.

The First of these fundamental effects is the Loss of Simultaneity. Here, two events that coincide in one frame (that is moving) do not coincide in another frame of reference. Take the example of a moving train, in which two flashlights placed at the center send photons to the opposite ends of the train. Within the train, they hit the ends at the same time, but because of the additional speed of the train, an outside observer would perceive the photons hitting the front of the train before the back end. This effect is paralleled by the next one, and we’ll go into the math with it.

This loss of simultaneity leads us to the Second Fundamental effect, Time Dilation. If you click on the link to see the post from September, the classic example of this effect is described by the observer on the train’s perception of light and the observer on the ground’s perception.

“The classic example Einstein gives in his papers on Special and General Relativity is the case of two persons, one on a train heading in a direction towards a point where a lightning bolt just hit the ground, and another person stands some distance away and can observe both the train and the lightning bolt.  In this scenario, the person on the train is moving at a speed v, who will hit the rays of light with this speed and thus (one would expect to) perceive the speed of photons as the intrinsic speed of light minus the speed of the person. This would differ from the person outside the train, seeing the lightning bolt come and thus (expect) to perceive light at its normal speed.

However, the speed of light must be CONSTANT at all times, so the person on the train must have something change in order for him to experience the speed of light at the correct value. What happens is, as Einstein describes in Special Relativity, that the person moving with speed v experiences time (and length and differently, it slows down in his (moving) reference frame and so the speed of light he perceives is maintained at the constant rate. For the person outside the train, he would see the light coming from the lightning bolt at its normal speed, without any special considerations to be taken into account. But as he sees the light going towards the train in the distance, he would clearly see the light in direction of the train has the same speed as the light that reached his reference frame, maintaining the constant speed of light and set consolidated with the perception of the person on the train, instead of two people experiencing two different speeds of light.”

Therefore, you can assume each person had a clock on them at the time of the lighting strike, and each perfectly measured the time at which they saw the light, and each would have different measurement. You can also picture this by seeing how two different paths light takes take different times to be completed, indicating that the longer path, the one where the frame itself was also moving. This is illustrated by:
                                 


As you can see, the initial case projects a different (and shorter) path once in the reference frame of A, than in the frame of B. So if the original distance was simply 2h, and the speed travelled c, so the time for the photons to traverse is \[\frac{2h}{c}.\] In the case of B, the total vertical distance the photons travel is 2h, but the speed component is different, being (by use of a bit of reasoning with the Pythagorean Theorem): \[\sqrt{c^2 – v^2},\] so the time would be: \[\frac{2h}{ \sqrt{c^2 – v^2}}.\] These two times can be compared and set as: \[t_A \sim t_B,\] \[ \frac{2h}{c} = \frac{2h}{ \sqrt{c^2 – v^2}},\] \[ 1 = \frac{c}{ \sqrt{c^2 – v^2}},\]\[ t_A = t_B \frac{1}{ \sqrt{1 – \frac{v^2}{c^2}}},\] where the additional factor for \(t_B\) is what is called \(\gamma_v\):\[\gamma_v = \frac{1}{\sqrt{1- \frac{v^2}{c^2}}},\] a common “conversion” factor which allows the switching of frames in Special Relativity. So the final time dilation equation is \[t_A = \gamma_v~ t_B, \] or \[t_B = \gamma_v~ t_A, \] depending on the reference from which the action is being observed.

This principle of time dilation, the factor \(\gamma_v\), establishes the base for most Relativistic problems (as long as they don’t involve gravity). 

The Third Fundamental Effect is Length Contraction. This effect establishes how an object moving at a speed is contracted proportionally to the increase in speed. This is described by \[L^\prime = \frac{L}{\gamma_v},\] describing how length is inversely proportional to speed.


These three fundamental effects are at the core of relativistic situations, which are then expanded upon by Einstein in 1915 with his treatise on General Relativity (100 Years of General Relativity has been floating around Science Media the last few months, if you hadn’t heard), where the impact of gravity is coalesced with relativity to describe myriad situations like Gravitational Lensing and Microlensing (see the posts about that by going back to the September group). These equations have become the basis for modern physics and astrophysics, forming part of the ongoing search to unite quantum mechanics and General Relativity, the greatest descriptions of the universe humans have ever had the ability to (sort of) prove. 

Just beyond the horizon, but no way we can see it

We are bounded in a nutshell of Infinite space: Blog Post #32, Worksheet # 10.1, Problem #3: Just beyond the horizon, but no way we can see it
3. Observable Universe. It is important to realize that in our Big Bang universe, at any given time, the size of the observable universe is finite. The limit of this observable part of the universe is called the horizon (or particle horizon to be precise, since there are other definitions of horizon.) In this problem, we’ll compute the horizon size in a matter dominated universe in co-moving coordinates. To compute the size of the horizon, let us compute how far the light can travel since the Big Bang.

(a) First of all - Why do we use the light to figure out the horizon size?

(b) Light satisfies the statement that \(ds^2 = 0\). Using the FRW metric, write down the differential equation that describes the path light takes. We call this path the geodesic for a photon. Choosing convenient coordinates in which the light travels in the radial direction so that we can set \(d\theta = d\phi = 0\), find the differential equation in terms of the coordinates t and r only.

(c) Suppose we consider a flat universe. Let’s consider a matter dominated universe so that \(a(t)\) as a function of time is known (in the last worksheet). Find the radius of the horizon today (at \(t= t_0\)).
(Hint: move all terms with variable r to the LHS and t to the RHS. Integrate both sides, namely r from 0 to \(r_{horizon}\) and t from 0 to \(t_0\).)

(a) As we all know, every modern theory of the large scale universe is based on several constants which are assumed and tested repeatedly to check for the consistency of the value. This is the case of the speed of light, c, the constant of the universe Einstein used to create special and general relativity, and the consequences of this knowledge have allowed us to understand the deeper throws of the universe. So, in order to find the radius of the horizon of the universe, we can use the only object which has reached us from the earliest point in time we can see, photons. Light has simply gotten to us before anything else, so it is the best representation of the horizon we are going to have.

(b) Simply put, we are going to find how the base FWR Metric: \[ds^2 = -cdt^2 + a^2(t) \left[\frac{dr^2}{1-kr^2} + r^2 (d\theta^2 + \sin^2\theta d\phi^2)\right],\]  can become a differential equation once we set \(ds^2\) to 0: \[ cdt^2 = a^2(t) \left[\frac{dr^2}{1-kr^2} + r^2 (d\theta^2 + \sin^2\theta d\phi^2)\right],\] \[ cdt^2 = \frac{a^2(t) (dr^2)}{1-kr^2},\] \[ \frac{cdt^2 }{a^2(t)} = \frac{ dr^2}{1-kr^2}.\] After all this rearrangement of the terms initially found, with both derivatives of angles set to 0, the equation becomes a relatively simple expression to describe the path light travels.


 (c) Furthermore, by establishing a flat universe, we have a definitive value for k, 0 in this case, so the equation becomes:\[ \frac{cdt^2 }{a^2(t)} = \frac{ dr^2}{1-kr^2},\]  \[ \frac{cdt^2 }{a^2(t)} = { dr^2},\]  which can now be set up to be solved in a way similar to the Friedmann Equations, using a parameter they established to complete the problem. So, from the Friedmann equation for a Matter Dominated Universe(http://ay17-rcordova.blogspot.com/2015/11/cosmology-101-part-2.html), the relation of the universal scale factor and time is: \[a(t) = a_0 \left(\frac{t}{t_0}\right)^{2/3} ,\] which can be placed into our derived form of the FRW metric and get: \[ dr^2= \frac{cdt^2 }{ a_0^2 \left(\frac{t}{t_0}\right)^{4/3}} ,\] which can be rewritten and analyzed in a manner similar to problem #2 of this Worksheet:  \[\int_0^{r_{horizon}} dr= \int_0^{t_0} \frac{cdt }{ a_0\left(\frac{t}{t_0}\right)^{2/3}} ,\] and after some integration and separation of variables, as well as recognizing \(a_0 \) and \(t_0\) as constants, we solve and get: \[ r_{horizon}= \frac{c t_0^{2/3}}{ a_0} \cdot 3t_0^{1/3},\] \[ r_{horizon}=3 \frac{c }{ a_0} t_0,\] which is the description for the radius of horizon today. 

Turns out Euclid was right about math after all

We are bounded in a nutshell of Infinite space: Blog Post #31, Worksheet # 10.1, Problem #2: Turns out Euclid was right about math after all

2. Ratio of circumference to radius. Let’s continue to study the difference between closed, flat and open geometries by computing the ratio between the circumference and radius of a circle.

(a) To compute the radius and circumference of a circle, we look at the spatial part of the metric and concentrate on the two-dimensional part by setting \(d\phi =0\)  because a circle encloses a two-dimensional surface. For the flat case, this part is just \[ds_{2d}^2= dr^2 + r^2d\theta^2.\]
The circumference is found by fixing the radial coordinate (\(r=\theta\) and \(dr=0\)) and both sides of the equation (note that \(\theta\) is integrated from 0 to \(2\pi\) ).
The radius is found by fixing the angular coordinate \((\theta , d\theta = 0 )\) and integrating both sides (note that dr is integrated from 0 to R).
Compute the circumference and radius to reproduce the famous Euclidean ratio \(2\pi\).

(b) For a closed geometry, we calculated the analogous two-dimensional part of the metric in Problem (1). This can be written as:\[ds_{2d}^2 = d\xi^2 + \sin\xi^2 d\theta^2.\]
Repeat the same calculation above and derive the ratio for the closed geometry. Compare your results to the flat (Euclidean) case; which ratio is larger? (You can try some arbitrary values of ξ to get some examples.)

(c) Repeat the same analyses for the open geometry, and comparing to the flat case.

(d) You may have noticed that, except for the flat case, this ratio is not a constant value. However, in both the open and closed case, there is a limit where the ratio approaches the flat case. Which limit is that?

(a) From the Friedman-Robertson-Walker Metric, an equation we will be looking on later in in this problem, we can derive a few basic equations that describes the physical phenomenon in universes with different configurations. Thought of either flat, open, or closed, the universe’s configuration describes how we perceive reality, where circles are the same in any location and their circumference is always \(2\pi R\), but this could be different for open or closed universes (even though astronomers observe the universe is most likely flat).
For a flat universe, we can prove the traditional Euclidean description of a circle with a few simple calculations: assuming \[d\phi = 0 ,\] we get: \[ds_{2d}^2= dr^2 + r^2d\theta^2.\]   This is a version of the FRW Metric with the \(d\phi\) set to 0.
Once we have this equation, we simply have to establish some parameters such as \(r = R, dr= 0\), and so the equation becomes: \[ds_{2d} = Rd\theta,\] , and by setting the integrals like the problem asks, we solve: \[\int_0^{Circumference}ds_{2d}=  \int_0^{2\pi}Rd\theta,\] \[Circumference = 2\pi R,\] which is exactly the traditional Euclidean case.
For the radius, we have to establish how \(\theta, d\theta= 0 \) and the equation now becomes: \[ds_{2d}^2= dr^2 + r^2d\theta^2,\] and following the same process as before for the circumference: \[ds_{2d}^2= dr^2,\]  \[\int_0^rds_{2d}= \int_0^R dr ,\] and so a true statement appears for the description of our universe: \[r= R.\]
Knowing the circumference and radius, we can use both of these to establish a ratio between them that becomes a standard ratio for comparison with other physical universe interpretations: \[ \frac{Circumference}{Radius} = \frac{2\pi R}{R} = 2\pi,\] which is the Euclidean model the problem asked for.

(b) For a closed universe, the equation is altered slightly because of the base FRW metric used, so we now have: \[ds_{2d}^2 = d\xi^2 + \sin\xi^2 d\theta^2,\] and with the alterations and limits the problem is establishing in order to better describe a closed universe, with  \[d\xi = 0 ,\] the equation now becomes \[ds_{2d} = \sin\xi d\theta.\] Next, using the hints the problem gives, we set up the integration: \[\int_0^{Circumference}ds_{2d}=  \int_0^{2\pi}\sin\xi d\theta,\]and now have \[Circumference = 2\pi \sin\xi.\]
Next we find the radius, with a process similar to the circumference, and the problem already set up how \(\theta, d\theta= 0 \), so the original equation now becomes: \[ds_{2d}^2 = d\xi^2 + \sin\xi^2 d\theta^2,\]   \[ds_{2d}^2= d\xi^2\] and integrating with the correct parameters: \[\int_0^rds_{2d}= \int_0^{\xi} d\xi ,\] we now have the radius: \[r= \xi.\]
Finally, establishing the ratio of circumference to ratio, we have: \[ \frac{Circumference}{Radius} = \frac{2\pi \sin\xi}{\xi}, \] which is identical to the flat universe ratio multiplied by the factor of \[\frac{\sin\xi}{\xi}. \]

(c) Taking the FRW metric as the basis, we can derive how: \[ds^2 = \frac{dr^2}{1-kr^2} + r^2 (d\theta^2 + \sin^2\theta d\phi^2),\] becomes \[ds^2 = \frac{dr^2}{1-kr^2} + r^2 d\theta^2,\] when \(d\phi = 0\). Then in the case of an open universe where k = -1, the equation becomes: \[ds^2 = \frac{dr^2}{1+r^2} + r^2 d\theta^2,\] so, taking a recommendation from an earlier problem of establishing that \(r = \sinh\xi \), the equation can now be solved to find a simpler solution which is similar to the other ratios. Now, the equation becomes: \[ds^2 = \frac{\cosh^2\xi \cdot d\xi^2}{1+\sinh^2\xi} + (\sinh\xi)^2 d\theta^2,\] and we can do this since an identity of hyperbolic functions like \(\sinh\xi\) establishes that its derivative is \(\cosh\xi d\xi\). Next, we simply use another identity says \(\cosh^2x – \sinh^2x = 1\), so the equation becomes: \[ds^2 = \frac{\cosh^2\xi \cdot d\xi^2}{\cosh^2\xi } + \sinh^2\xi d\theta^2,\] \[ds^2 = d\xi^2 + \sinh^2\xi d\theta^2. \] Using the exact same process as earlier in part b, we find that the ratio of circumference to radius is: \[\frac{Circumference}{Radius}=\frac{2\pi \sinh\xi}{\xi},\] which is the same as the flat universe ratio multiplied by the factor of \[\frac{\sinh\xi}{\xi}. \]

(d) From these ratios, a clear patter can be discerned once compared to the original, flat universe, case. With all the ratios illustrated clearly: (flat, closed, and open, respectively) \[2\pi , ~ \frac{2\pi \sin\xi}{\xi} , ~ \frac{2\pi \sinh\xi}{\xi}\] there is a clear change. Both open and closed universe models exhibit the trigonometric and hyperbolic function, which both happen to have a similar Taylor Expansion (a way to approximate for small numbers), which turns the ratios into: \[2\pi , ~ \frac{2\pi \xi}{\xi} , ~ \frac{2\pi \xi}{\xi}\] \[2\pi , ~2\pi , ~ 2\pi.\] Therefore, as \(\xi\) becomes small, all universes look more and more like each other, as a result of the limiting case as \[\xi \to 0 .\]

Monday, November 9, 2015

Telescopes in the night, but actually used in the day

We are bounded in a nutshell of Infinite space: Blog Post #30, Worksheet # 9.2: Telescopes in the night, but actually used in the day

1. Noise properties of a single dish radio telescope

Radio telescopes (or radiometers) work like a standard radio antenna or satellite dish, converting an incoming radio signal into a pattern of oscillating voltages through the interaction of the wave with the electrons in the antenna. A radiometer receives a large number of signals, but filters allow only a certain chunk of the spectrum through. Our telescope is what has called a “heterodyne” receiver, which means that it takes the incoming signal at \(\nu_0\), and mixes it down with a local oscillator of frequency \(\nu_{LO}\) to produce a signal centered at a lower new frequency \(\nu_{IF} = \nu_0 - \nu_{LO}\). The system is designed so that νif is a constant. So, when we “tune” the telescope, we are really setting \(\nu_{LO}\). This signal is then fed into the filter bank and split into \(N_{chan}\)  channels of width \(\Delta\nu_{RF} \). The power in each channel is plotted, and this is called a spectrum. Note that \(T_{sys}\) is called the “system temperature” and has units of Kelvin; however, you should think of \(T_{sys}\) as a measurement of noise and not a physical temperature.
When doing this, the signal has very particular noise properties. Recall that the root-mean-square (rms) error decreases as a function of the number of samples \( \sigma \sim 1/\sqrt{N}\). For spectral line observations like ours, in a radiometer, each sample has noise \(T_{sys}\), and the number of samples in a channel goes up with increasing channel width and integration time. The noise is characterized by the ideal radiometer equation.
Precisely, if I have a receiver with a system temperature \(T_{sys}\), channel width \(\Delta\nu\), and integration time \(\tau\) , the radiometer equation is \[\sigma = \frac{T_{sys}}{\sqrt{\tau ~\Delta \nu}}.\]

(a) What is the expression for the signal to noise ratio (SNR)? The signal has a brightness temperature \(T_{A}\).

(b) Find an equation for the velocity resolution \(\Delta\nu\) km/s that corresponds to a channel width of \(\Delta\nu\) MHz at the frequency of \(^{12}CO, \nu =\)115.271 GHz (2.6 mm). Write this in the form \[\Delta\nu = \_\_\_\_ km/s \left(\frac{\Delta\nu}{1~MHz}\right) \]  

(c) What integration time would be needed to detect the peak of \(^{12}CO\) with SNR = 10 if I use a filter bank with 256 channels that are 0.5 MHz wide. \(T_{A}\) for CO is about 2-3 K and \(T_{sys} = 500 K\). (Hint: The peak appears in 1 channel. Do the algebra first to get the final expression and then put in the numbers.)

2. Resolution of a single dish radio telescope
The spatial resolution of a telescope is \(\theta = 1.2\frac{\lambda}{D}\) , where D is the diameter of the dish, and \(\lambda\) is the observing wavelength. In radio astronomy, \(\theta\) is also known as the half-power beam width (or full-width half-max of the beam).

(a) Find an equation for the beam width, in arcminutes, of a single-dish radio in terms of frequency \(\nu\) in GHz, and diameter D in meters. Use a calculator to determine this to 1 decimal place. Write an equation of the form \[\theta_{HPBW} = \_\_\_\_ degree \left(\frac{\nu}{GHz}\right)^{-1} \left(\frac{D}{m}\right)^{-1} \]

(b) What is the beamwidth, in arcminutes, for the CfA 1.2 m telescope at the \(^{12}CO\) frequency. (to 1 decimal place)

(c) What linear dimension in pc does this correspond to at the Galactic center (8.5 kpc)?

3. Geometry of the experiment
Below is a schematic of the experiment. A molecular cloud a distance R from the galactic center has a velocity \(V_{cir}\). When we observe this cloud, we detect a different velocity, \(V_{obs}\), which is the combination of our velocity towards it and its velocity away. Note: We are using R to denote the distances because the sources are some radius from the Galactic center.


(a) For our experiment we will take a spectra which will have velocity components from multiple clouds along the line of sight. Looking at the figure above, which cloud will have the highest velocity and why?

(b) Plot, very roughly, the velocity of the clouds versus their distance along the line sight, what would the shape be?

(c) The velocity which we measure using the \(^{12}CO\) line is the combination of the cloud’s velocity around the galactic center and the projection of the Sun’s velocity along our line of sight. What is the equation for the velocity we measure \(V_{max}\)?

(d) Write the equations for the rotational velocity, \(V_{cir}\)and \(R_{tan}\)in terms of the quantities we know - \(V_{\odot}\), \(V_{\odot}\), l, and \(V_{max}\) .

(e) What is the equation for the mass profile of the galaxy? The ultimate goals of this lab are to estimate the total gravitational mass of the Galaxy within the Suns orbit and to infer something about how that mass is distributed in Galactic radius. To do this we will determine the so-called Galactic rotation curve by measuring the velocities of molecular clouds orbiting in the Galactic potential.

1. (a) For this post, we will be exploring the nature of a radio telescope and how these can be used and their slight limits corrected by knowing exactly what to look for and what effects to mitigate. First of all, we’ll find the ratio of the signal to the noise. So if the noise has already been defined as \[\sigma = \frac{T_{sys}}{\sqrt{\tau ~\Delta \nu}},\] and the signal is defined as \(T_A\), then it is simply a matter of dividing one by the other. So the Signal to Noise Ratio is simply: \[\frac{T_A}{\sigma}\] \[\frac{T_A}{T_{sys}} \sqrt{\tau~\Delta\nu}.\]

(b) Next, we are asked to find the relationship between velocity and frequency, something that sounds very familiar to red shift. In fact, from the red shift equation, with a bit changed, we have: \[\frac{\nu_{obs} - \nu_{em} }{\nu_{em}} = \frac{v}{c} , \] which can be interpreted as: \[\Delta v = c\frac{\Delta\nu}{\nu}.\] So taking this equation and plugging in the necessary values as seen above, we have: \[ \Delta v = 3 \times 10^5 km/s \cdot \frac{\Delta\nu}{115.271 \times 10^9 Hz} ,\] which yields: \[\Delta v = 2.6 \times 10^{-6} \Delta\nu \frac{km/s}{Hz} \times \frac{10^6 Hz}{MHz},\] adding in the conversion factor to achieve the correct units as asked, we finally know that: \[\Delta v = 2.6 \frac{km/s}{MHz} \Delta\nu .\]

(c) Now, knowing how the velocity resolution equation works and how the SNR operates, we can use them to find the next factor, the time of integration for \(^{12} CO\), \(\tau\). Using the SNR equation, we have:   \[SNR = \frac{T_A}{T_{sys}} \sqrt{\tau~\Delta\nu}, \] and solving for the integration time, \[\tau =\frac{SNR^2\frac{T_{sys}^2}{T_A^2}}{\Delta \nu} .\] By plugging in the values given to us by the question, we have: \[\tau = \frac{10^2\frac{500K^2}{2.5K^2}}{0.5 \times 10^6 Hz},\] which simplifies down to: \[\tau = 8 s.\]

2.(a) Here, we are given a basic piece of information that essentially tells us what to do in order to find the final angle. Taking into consideration the \(\theta = 1.2\frac{\lambda}{D}\), it can be turned into \(\theta = 1.2\frac{\frac{c}{\nu}}{D}\), so by plugging this into the equation given, we can make it in the correct dimensions by multiplying by the correct orders of magnitude. So: \[\theta_{HPBW} = \_\_\_\_ degree \left(\frac{\nu}{GHz}\right)^{-1} \left(\frac{D}{m}\right)^{-1} ,\] \[\theta_{HPBW} = 1.2\frac{\frac{c}{\nu}}{D} \left(\frac{\nu}{GHz}\right)^{-1} \left(\frac{D}{m}\right)^{-1} ,\]  \[\theta_{HPBW} = \frac{1.2 \cdot 3\times 10^8 m/s }{ \nu D} \times \frac{1~GHz}{10^{-9} Hz}  \times \frac{180~degrees}{\pi ~radians} ,\] which it all simplifies to: \[\theta_{HPBW} = \frac{20.36~degrees~ m\cdot GHz}{\nu~D} . \]

(b) Now we take the equation we just found and plug in the values of D and \(\nu\) they give. So the equation turns into: \[\theta_{HPBW} = \frac{20.36~degrees}{\nu~D} , \] \[\theta_{HPBW} = \frac{20.36~degrees~ m\cdot GHz}{1.2 m \cdot 115.271 GHz}, \] \[\theta_{HPBW} = 0.149 ~degrees = 8.9 ~arcminutes.\]

(c) Now that we know the angle with which we are observing, we can use it in the basic angle to distance relation that is parallax. From the information given by the question, we know that the distance from the object is 8.5 kpc, and if we have the distance and the angle, then we can know the width of our field of view. From earlier problems, we found how \[angle_{(in~ arcseconds)} =\frac{1 AU}{D_{in ~pc}}, \] but the relationship can be rewritten as: \[angle_{(in~ arcseconds)} \cdot D_{(in~pc)} = width ~of ~view ~in~ AU , \] and we know the values that need to be plugged in and the conversion factors to be used: \[(8.9 ~arcminutes \times 60) \cdot (8.5 \times 10^3 pc) = width ,\] \[ 4.539 \times 10^6 AU  \ times \frac{1 ~pc }{2.05 \times 10^5 AU}= width,\] \[Linear~Dimension_{(Width)} = 22.14 pc. \]

3. (a) After looking at the diagram for a bit, it becomes obvious that the object with the highest speed away from us, the object with the greatest Red Shift, is undoubtedly object B, since all the velocity we observe is pointed tangentially to its orbit path, thus making it the fastest among all the objects in our line of sight.

(b)


(c) Taking into consideration all the factors that are changing how we are perceiving the objects on the other side of the galaxy,  the best equation that relates the max speed of any object is: \[V_{Max} = V_{cir} - V_\odot \sin(l) , \] which recognizes how our velocity impacts the way we perceive the velocity of the observed object.

(d) By simply rewriting the above equation, we can solve for the circular velocity, which is simply: \[ V_{cir} = V_{Max} + V_\odot \sin(l) , \] and the tangential radius is simply the component of the radius from the galactic core to the sun, such that: \[R_{tan} = R_\odot \sin(l) .\]


(e) And finally, to understand the mass profile for the galaxy, we must go back to a post from a couple of months ago ( http://ay17-rcordova.blogspot.com/2015/09/the-milky-way-is-now-sphere-trust-me.html ) where the velocity profile is described as: \[v(r) = \left(\frac{G M_{enc}}{r}\right)^{1/2} , \] and to find the mass profile, a bit of rearranging is all that is necessary: \[ M_{enc} = \frac{V_{cir}^2 R_{tan}}{G}.\]  

Cosmology 101 Part 2

We are bounded in a nutshell of Infinite space: Blog Post #29, Worksheet # 9.1, Problem #2: Cosmology 101 Part 2

2. GR modification to Newtonian Friedmann Equation:
In Question 1, you have derived the Friedmann Equation in a matter-only universe in the Newtonian approach. That is, you now have an equation that describes the rate of change of the size of the universe, should the universe be made of matter (this includes stars, gas, and dark matter) and nothing else. Of course, the universe is not quite so simple. In this question we’ll introduce the full Friedmann equation which describes a universe that contains matter, radiation and/or dark energy. We will also see some correction terms to the Newtonian derivation.

(a) The full Friedmann equations follow from Einstein’s GR, which we will not go through in this course. Analogous to the equations that we derived in Question 1, the full Friedmann equations express the expansion/contraction rate of the scale factor of the universe in terms of the properties of the content in the universe, such as the density, pressure and cosmological constant. We will directly quote the equations below and study some important consequences. The first Friedmann equation:
\[ \left(\frac{\dot{a}}{a}\right)^2 = \frac{8\pi}{3}G\rho - \frac{kc^2}{a^2} + \frac{\Lambda}{3}. \]
The second Friedmann equation:
\[ \frac{\ddot{a}}{a} = -\frac{4\pi G }{3c^2} (\rho c^2 + 3P) + \frac{\Lambda}{3}.\]
In these equations, \(\rho\) and P are the density and pressure of the content, respectively. k is the curvature parameter; k = -1, 0, 1 for open, flat and closed universe, respectively. \(\Lambda\) is the cosmological constant. Note that in GR, not only density but also pressure are the sources of energy. Starting from these two equations, derive the third Friedmann equation, which governs the way average density in the universe changes with time.
\[\dot{\rho} c^2 = -3 \frac{\dot{a}}{a}(\rho c^2 + P). \]   To derive this equation, first multiply \(a^2\)  on both sides of  the first equation and then take time derivative on both sides; plug the second equation into your expression to eliminate \(\ddot{a}\).
Now we can use these equations to derive some fun consequences of different kinds of universe, some of which describe our own universe. For simplicity, in the exercise below, let us always set \(k = 0 \) , namely consider a flat universe. Luckily for us, state-of-the-art observations suggest that our universe is likely flat.

(b) Cold matter dominated universe. If the matter is cold, its pressure P = 0, and the cosmological constant \(\Lambda\). Use the third Friedmann equation to derive the evolution of the density of the matter \(\rho\) as a function of the scale factor of the universe a. You can leave this equation in terms of \(\rho , \rho_0 , a \) and \(a_0\) , where \(\rho_0\)  and \(a_0\)  are current values of the mass density and scale factor. The result you got has the following simple interpretation. The cold matter behaves like “cosmological dust” and it is pressureless (not to be confused with warm/hot dust in the interstellar medium!). As the universe expands, the mass of each dust particle is fixed, but the number density of the dust is diluted - inversely proportional to the volume.
Using the relation between \(\rho\)  and a that you just derived and the first Friedmann equation, derive the differential equation for the scale factor a for the matter dominated universe. Solve this differentiation equation to show that \(a(t) \propto t^{2/3}\)  . This is the characteristic expansion history of the universe if it is dominated by matter. (Hint: When solving this final differential equation, recall that at time t = 0, a = 0 (the Big Bang). At time \(t = t_0 , a = a_0 = 1\)  (present day).)

(c) Radiation dominated universe. Let us repeat the above exercise for a universe filled with radiation only. For radiation, \(P = \frac{1}{3} \rho c^2 \)  and \(\Lambda = 0\). Again, use the third Friedmann equation to see how the density of the radiation changes as a function of scale factor.
The result also has a simple interpretation. Imagine the radiation being a collection of photons. Similar to the matter case, the number density of the photon is diluted, inversely proportional to the volume. Now the difference is that, in contrast to the dust particle, each photon can be thought of as wave. As you learned last week, the wavelength of the photon is also stretched as the universe expands, proportional to the scale factor of the universe. According to quantum mechanics, the energy of each photon is inversely proportional to its wavelength: \(E = h\nu\). Unlike the dust case where each particle has a fixed energy. So in an expanding universe, the energy of each photon is decreasing inversely proportional to the scale factor. Check that this understanding is consistent with the result you got. Again using the relation between \(\rho\) and a and the first Friedmann equation to show that \(a(t) \propto t^{1/2}\)  for the radiation only universe.

(d) Cosmological constant/dark energy dominated universe. Imagine a universe dominated by the cosmological-constant-like term. Namely in the Friedmann equation, we can set \(\rho = 0\)  and P ­= 0 and only keep \(\Lambda\) nonzero. As a digression, notice that we said “cosmological-constant-like” term. This is because the effect of the cosmological constant may be mimicked by a special content of the universe which has a negative pressure \( P = - \rho c^2\)  . Check that the effect of this content on the right-hand-side of third Friedmann equation is exactly like that of the cosmological constant. To be general we call this content the Dark Energy. How does the energy density of the dark energy change in time? Show that the scale factor of the cosmological-constant-dominated universe expands exponentially in time. What is the Hubble parameter (constant) of this universe?
Hint: While calculating the scale factor as a function of time, you will find that setting \(a(0) = 0\) leads to a negative infinity. Feel free to ignore this term to show the dependence.

(e) Suppose the energy density of a universe at its very early time is dominated by half matter and half radiation. (This is in fact the case for our universe 13.7 billion years ago and only 60 thousand years after the Big Bang.) As the universe keeps expanding, which content, radiation or matter, will become the dominant component? Why?

(f) Suppose the energy density of a universe is dominated by similar amount of matter and dark energy. (This is the case for our universe today. Today our universe is roughly 68% in dark energy and 32% in matter, including 28% dark matter and 5% usual matter, which is why it is accelerated expanding today.) As the universe keeps expanding, which content, matter or the dark energy, will become the dominant component? Why? What is the fate of our universe?

(a) Following the above instructions in order to derive the third Friedmann equation, we first take the first equation: \[ \left(\frac{\dot{a}}{a}\right)^2 = \frac{8\pi}{3}G\rho - \frac{kc^2}{a^2} + \frac{\Lambda}{3}, \] and multiply both sides by the scale factor a squared. \[  \dot{a}^2 = \frac{8\pi}{3}G\rho a^2 - kc^2 + \frac{\Lambda}{3} a^2, \] and now taking the time derivative, with respect to the density \(\rho\) and the scale factor a: \[2\dot{a} \ddot{a} = \frac{8\pi}{3}G(a^2 \dot{\rho} + 2\rho a \dot{a}) + \frac{\Lambda}{3} \cdot 2 a \dot{a}, \] and solving for \(\ddot{a}\) : \[\ddot{a} = \frac{\frac{4\pi}{3}G}{\dot{a}}(a^2 \dot{\rho} + 2\rho a \dot{a}) + \frac{\Lambda}{3}\cdot a,\] \[\ddot{a} = \frac{ 4\pi G a^2 \dot{\rho}}{3 \dot{a}} + \frac{8 \pi G \rho a}{3} + \frac{\Lambda}{3}\cdot a,\] and now return one scale factor a to the denominator of the left side of the equation,  \[\frac{\ddot{a}}{a} = \frac{ 4\pi G a \dot{\rho}}{3 \dot{a}} + \frac{8 \pi G \rho }{3} + \frac{\Lambda}{3},\] and now this can be equated to the Second Friedmann equation, seeing as they both have \(\frac{\ddot{a}}{a}\) as one side of their equation. So:
\[\frac{\ddot{a}}{a} = \frac{ 4\pi G a \dot{\rho}}{3 \dot{a}} + \frac{8 \pi G \rho }{3} + \frac{\Lambda}{3},\]
\[ \frac{\ddot{a}}{a} = -\frac{4\pi G }{3c^2} (\rho c^2 + 3P) + \frac{\Lambda}{3}.\]
\[\frac{ 4\pi G a \dot{\rho}}{3 \dot{a}} + \frac{8 \pi G \rho }{3} + \frac{\Lambda}{3} =  -\frac{4\pi G }{3c^2} (\rho c^2 + 3P) + \frac{\Lambda}{3},\] which allows us to do some quick algebra to cancel out al lot of terms, \[\frac{ 4\pi G a \dot{\rho}}{3 \dot{a}} + \frac{8 \pi G \rho }{3} =  -\frac{4\pi G }{3c^2} (\rho c^2 + 3P) ,\] \[\frac{ \pi G a \dot{\rho}}{ \dot{a}} + 2 \pi G \rho  =  -\frac{\pi G }{c^2} (\rho c^2 + 3P) ,\]  \[\frac{ a \dot{\rho}}{ \dot{a}} + 2 \rho  = -\rho - \frac{3P}{c^2} ,\] \[\dot{\rho}\frac{ a }{ \dot{a}} = -3\rho - \frac{3P}{c^2} ,\] \[\dot{\rho}\frac{ a }{ \dot{a}} c^2 = -3\rho c^2 - 3P ,\] \[\dot{\rho} c^2 = -3\frac{\dot{a}}{ a }(\rho c^2 + P),\] which is the third Friedmann equation we have been trying to solve for.

(b) In the case of a cold matter dominated universe, there are several considerations that can and should be taken advantage of, which are all outlined by the instructions. These conditions (\(\Lambda = 0 \), P = 0 and k = 0 ), turn the third equation into: \[\dot{\rho} c^2 = -3\frac{\dot{a}}{ a }(\rho c^2),\] \[\frac{\dot{\rho}}{\rho} = -3\frac{\dot{a}}{ a },\] and we take the antiderivative with respect to time on both sides: \[ \int_{\rho_0} ^\rho \frac{\frac{d\rho}{dt}}{\rho} = -3 \int_{a_0} ^a \frac{\frac{da}{dt}}{ a },\] which turns into: \[\ln(\frac{\rho}{\rho_0} )= -3 \ln(\frac{a}{a_0}) , \] and can be simplified into: \[\frac{\rho}{\rho_0} = \frac{a_0 ^3}{a^3}\ ,\] \[\rho = \frac{\rho_0 a_0 ^3}{a^3} .\] And since the question also states that  \(t = t_0 , a = a_0 = 1\)  and \(\rho_0 = 1\), so the relationship becomes: \[\rho = \frac{1}{a^3} .\]
We now take this equation and combine it with the First Friedmann equation, and get: \[ \left(\frac{\dot{a}}{a}\right)^2 = \frac{8\pi}{3}G\rho ,\] \[ \left(\frac{\dot{a}}{a}\right)^2 = \frac{8\pi}{3}G \frac{1}{a^3} ,\]  \[ \dot{a}^2 = \frac{8\pi}{3}G \frac{1}{a} ,\] \[ \dot{a} = a^{-1/2}\sqrt{\frac{8\pi}{3}G } ,\] and since we are only looking for a proportion, the constants can be ignored and considered c: \[ \dot{a} = a^{-1/2} c ,\] now we take an integral on both sides: \[ \frac{da}{dt} = a^{-1/2} c ,\] \[ a^{1/2}da = c \cdot dt ,\] \[ \int_0 ^a a^{1/2}da =\int_0 ^t c \cdot dt ,\]  \[\frac{2}{3} a^{3/2} = t c,\] \[ a = t^{2/3} \frac{3}{2}c,\] therefore: \[ a(t) \propto t^{2/3} . \] This is the characteristic expansion history of the universe if it is dominated by matter.

(c) Another case for the development of the universe is the way it might be radiation dominated, meaning the entirety of the electromagnetic spectrum. Here, we follow a very similar procedure to the last problem, except now we take \(P = \frac{1}{3} \rho c^2\), which turns the equation into: \[\dot{\rho} c^2 = -3\frac{\dot{a}}{ a }(\rho c^2 + \frac{1}{3} \rho c^2),\] \[\dot{\rho} c^2 = -4\frac{\dot{a}}{ a }(\rho c^2),\] which makes a small but meaningful difference once we follow the same process of integration we saw in the previous problem. The integration leaves us with: \[\rho = \frac{1}{a^4} \], which is now plugged into the First Friedmann equation, and we have: \[ \left(\frac{\dot{a}}{a}\right)^2 = \frac{8\pi}{3}G \frac{1}{a^4} ,\] and once again following the exact same process of separation of variables, we have: \[ \frac{da}{dt} = a^{-1} c ,\] which can be evaluated by integrating, and we get: \[ a(t) \propto t^{1/2} ,\] which describes the development of a radiation only universe.

(d) For this problem, different considerations have to be taken into account, mainly the factor we have been eliminating in the other problems, \(\Lambda\). Here, all the terms with \(\rho\) and P will cancel out and we will be left with, from the First Friedmann equation: \[ \left(\frac{\dot{a}}{a}\right)^2 =  \frac{\Lambda}{3}, \] \[ \frac{\dot{a}}{a} =  \sqrt{\frac{\Lambda}{3}}, \]which can be expanded to find a general equation relating the development rate to the new Hubble Parameter. So we have: \[ \frac{\frac{da}{dt}}{a} =  \sqrt{\frac{\Lambda}{3}}, \] \[ \frac{da}{a} = \sqrt{\frac{\Lambda}{3}} dt , \] \[\int_{a_0} ^a \frac{da}{a} = \int_{t_0} ^t \sqrt{\frac{\Lambda}{3}} dt , \] and keeping the definitions we had for initial time and expansion rate, we get:  \[ \ln(a) =  \sqrt{\frac{\Lambda}{3}} t , \]  \[ a =  e^{\sqrt{\frac{\Lambda}{3}} t} = exp\left(\sqrt{\frac{\Lambda}{3}} t \right) .\] So we now know the Hubble parameter of the universe is \(\sqrt{\frac{\Lambda}{3}}\), and that the expansion of this dark Energy dominated Universe is exponential.

(e) As the universe continues expanding, matter would become the dominating component since its rate of expansion is larger and its slope of separation (-3 vs. -4 for radiation) is lower so it will continue to be present and dominate after the radiation has dissipated.

(f) Dark energy will become the dominant component in the universe, since its expansion is the greatest of all, since it depends on a true exponential increase at all times. Eventually, the universe would be so expanded, that gravity no longer would exert enough force to pull matter back in on itself, disproving the “Big Crunch” Theory, and the dark energy would continue propagating as far as possible, turning the universe into a cold and seemingly barren place, with hundreds of orders of magnitudes of space for every speck of matter.  



Cosmology 101 Part 1

We are bounded in a nutshell of Infinite space: Blog Post #28, Worksheet # 9.1, Problem #1: Cosmology 101 Part 1

1. A Matter-only Model of the Universe in Newtonian approach
In this exercise, we will derive the first and second Friedmann equations of a homogeneous, isotropic and matter-only universe. We use the Newtonian approach. Consider a universe filled with matter which has a mass density \(\rho(t)\) . Note that as the universe expands or contracts, the density of the matter changes with time, which is why it is a function of time t.
Now consider a mass shell of radius R within this universe. The total mass of the matter enclosed by this shell is M. In the case we consider (homogeneous and isotropic universe), there is no shell crossing, so M is a constant.

(a) What is the acceleration of this shell? Express the acceleration as the time derivative of velocity, \(\dot{v}\) (pronounced v-dot) to avoid confusion with the scale factor a (which you learned about last week).

(b) To derive an energy equation, it is a common trick to multiply both sides of your acceleration equation by v. Turn your velocity, v, into \(\frac{dR}{dt}\) , cancel dt, and integrate both sides of your equation. This is an indefinite integral, so you will have a constant of integration; combine these integration constants call their sum C. You should arrive at the following equation  \[\frac{1}{2}\dot{R}^2 - \frac{GM}{R} = C\] Convince yourself the equation you’ve written down has units of energy per unit mass.

(c) Express the total mass M using the mass density, and plug it into the above equation. Rearrange your equation to give an expression for \(\left(\frac{\dot{R}}{R}\right)^2 \) , where \(\dot{R}\) is equal to \(\frac{dR}{dt}\)  .

(d) R is the physical radius of the sphere. It is often convenient to express R as \(R = a(t)r\) , where r is the comoving radius of the sphere. The comoving coordinate for a fixed shell remains constant in time. The time dependence of R is captured by the scale factor \(a(t) \) . The comoving radius equals to the physical radius at the epoch when \(a(t) = 1\) . Rewrite your equation in terms of the comoving radius, r, and the scale factor, \(a(t)\) .

(e) Rewrite the above expression so that \(\left(\frac{\dot{a}}{a}\right)^2\) appears alone on the left side of the equation.

(f) Derive the first Friedmann Equation: From the previous worksheet, we know that \(H(t) =\frac{\dot{a}}{a}\) . Plugging this relation into your above result and identifying the constant \(2C/r^2 = -kc^2\)  where k is the “curvature” parameter, you will get the first Friedmann equation. The Friedmann equation tells us about how the shell of expands or contracts; in other words, it tells us about the Hubble expansion (or contracration) rate of the universe.

(g) Derive the second Friedmann Equation: Now express the acceleration of the shell in terms of the density of the universe, and replace R with \(R = a(t)r\). You should see that \(\frac{\ddot{a}}{a}= -\frac{4\pi}{3}G\rho \) , which is known as the second Friedmann equation.
The more complete second Friedmann equation actually has another term involving the pressure following from Einstein’s general relativity (GR), which is not captured in the Newtonian derivation.
If the matter is cold, its pressure is zero. Otherwise, if it is warm or hot, we will need to consider the effect of the pressure.

(a) In any frame, there are basic equations that permit us to physically interpret the system. One is: \[ F = ma\] as per Newton’s second law of motion. This can be rewritten as \[F= m\dot{v}\] since the dot represents the derivative of the variable below it. Now with this, the force of gravity can be equated with it: \[F_g = \frac{GMM}{R^2} \] \[ m\dot{v}=- \frac{GMM}{R^2}\] and canceling out the appropriate terms, we’re left with \[\dot{v}=-\frac{GM}{R^2},\] the acceleration of the shell.

(b) Multiplying the above expression by v  on both sides, we have: \[v \cdot\dot{v}=-\frac{GM}{R^2} \cdot v ,\] which can be rewritten as: \[v \frac{dv}{dt}=-\frac{GM}{R^2} \frac{dR}{dt},\] and canceling out the dt’s and integrating both sides with respect to the correct limits, we have: \[\frac{v^2}{2} = \frac{GM}{R} + C ,  \] which is  \[\frac{1}{2}\dot{R}^2 - \frac{GM}{R} = C\] once rearranged.

(c) The total mass of this system can be expressed as the density multiplied by volume, such that: \[M = \rho V \] and V can also be rewritten for a spherical system as: \[ V = \frac{4}{3} \pi R^3 , \] and plugging both of these into the above equation, we get:  \[\frac{1}{2}\dot{R}^2 - \frac{GM}{R} = C, \] \[\frac{1}{2}\dot{R}^2 - \frac{G\rho\frac{4}{3} \pi R^3 }{R} = C, \] and by working with the algebra a bit, and taking into consideration that the C is a fixed value, we have: \[\frac{1}{2}\dot{R}^2  =G\rho\frac{4}{3} \pi R^2  + C, \]  \[\dot{R}^2  =G\rho\frac{8}{3} \pi R^2 + 2C, \] \[\left(\frac{\dot{R}}{R}\right)^2  =G\rho\frac{8}{3} \pi  + \frac{2C}{R^2}, \] which is what the question asked for.

(d) Now we simply rewrite the equation with a slight change to redefine R as something else. Here, \(R = a(t)r\), and as such: \[\left(\frac{\dot{a}(t)r}{a(t)r}\right)^2  =G\rho\frac{8}{3} \pi  + \frac{2C}{(a(t)r)^2}, \] what we do next is a bit more interesting.

(e) Now, we simplify the expression back into a more manageable system with the right side that we want: \[\left(\frac{\dot{a}}{a}\right)^2  =G\rho\frac{8}{3} \pi  + \frac{2C}{a(t)^2 r^2}. \]

(f) Next, it is simply a matter of following the instructions correctly, just substituting \( \frac{\dot{a}}{a}\) with H(t) and \(2C/r^2\)  with \( -kc^2\). So we have: \[H(t)^2  =\frac{8}{3}\pi G\rho - \frac{kc^2}{a^2}, \] which is the Newtonian mechanics version of the first Friedmann equation.


(g) For the Second Friedmann equation, it is just a matter of going back to the beginning of the previous derivation process we just did, and having: \[\dot{v}=-\frac{GM}{R^2},\] and at this moment plugging in the definition of mass we had established previously, \[ M =\rho \frac{4}{3} \pi R^3 , \] we know simplify down to: \[\dot{v}=-\frac{G\rho \frac{4}{3} \pi R^3 }{R^2},\]   \[\dot{v}= -\frac{4}{3} \pi G\rho R,\] and redefining R and a as we did beforehand, \[\ddot{a} (t) r = -\frac{4}{3} \pi G\rho a(t) r,\] and placing everything and simplifying a bit more, we get:  \[\frac{\ddot{a}}{a} = -\frac{4}{3} \pi G\rho,\] the other Newtonian mechanics version of the Friedmann equations. 

Monday, November 2, 2015

The universe is REALLY old and REALLY big

We are bounded in a nutshell of Infinite space: Blog Post #27, Worksheet # 8.1, Problem #3: The universe is REALLY old and REALLY big

3. It is not strictly correct to associate this ubiquitous distance-dependent redshift we observe with the velocity of the galaxies (at very large separations, Hubble’s Law gives ‘velocities’ that exceeds the speed of light and becomes poorly defined). What we have measured is the cosmological redshift, which is actually due to the overall expansion of the universe itself. This phenomenon is dubbed the Hubble Flow, and it is due to space itself being stretched in an expanding universe.

Since everything seems to be getting away from us, you might be tempted to imagine we are located at the center of this expansion. But, as you explored in the opening thought experiment, in actuality, everything is rushing away from everything else, everywhere in the universe, in the same way. So, an alien astronomer observing the motion of galaxies in its locality would arrive at the same conclusions we do.

In cosmology, the scale factor,\(a(t)\)  is a dimensionless parameter that characterizes the size of the universe and the amount of space in between grid points in the universe at time t. In the current epoch, \( t = t_0\)  and \(a(t_0) \equiv 1\)  . \(a(t)\)  is a function of time. It changes over time, and it was smaller in the past (since the universe is expanding). This means that two galaxies in the Hubble Flow separated by distance \(d_0 = d(t_0) \) in the present were \(d(t) = a(t) d_0 \)   apart at time t.

The Hubble Constant is also a function of time, and is defined so as to characterize the fractional rate of change of the scale factor:

\[H(t) = \frac{1}{a(t)} \frac{da}{dt}|_t\] and the Hubble Law is locally valid for any t: \[ v = H(t)d\] where v is the relative recessional velocity between two points and d the distance that separates them.

(a) Assume the rate of expansion, \(\dot{a} = da/dt \), has been constant for all time. How long ago was the Big Bang (i.e. when \(a(t=0) = 0\) )? How does this compare with the age of the oldest globular clusters\(\sim 12 Gyr)? What you will calculate is known as the Hubble Time.

(b) What is the size of the observable universe? What you will calculate is known as the Hubble Length.

(a) In this problem, we’ll be examining some of the first methods developed that correctly estimated the age, and size of the universe (from our perspective). Going back to one of the pioneers in the field of cosmology, Edwin Hubble established a set of relationships and constants which have largely determined the basis for all cosmology. One of these basics is the Hubble constants, a value found by analyzing the Red Shifts of many galaxies and comparing them all in their velocities and distance from Earth. The slope of this graphical comparison yields the Hubble Constant, which can be approximated to being \( 70 \frac{\frac{km}{s}}{Mpc} \). Knowing this, we can start to use calculus and integrate the above calculation for the Hubble Constant as a function of time when \(a(t) = a(t_0)\): \[H(t_0) = \frac{1}{a(t_0)} \frac{da}{dt}|_{t=0} , \] \[H(t_0) = \frac{da}{dt} , \] \[ dt = \frac{da}{ H(t) } , \] and now we integrate to find the total \(H(t)\) : \[\int_{t=0}^{t=present} dt = \int^{a(t_0)}_0 \frac{da}{ H(t) } , \] \[t = H_0 ^{-1} .\] And now, plugging in the compiled value for Hubble Time, we can convert it to an actual time by making the dimensional analysis work. So if \( H_0 = 70 \frac{\frac{km}{s}}{Mpc} , \) then we simply find the value of km in pc, and turn Mpc into pc, so the final dimensional analysis would be: \[ H_0 = 70 \frac{\frac{km}{s}}{Mpc} \cdot ( 3.24 \times 10^{-20} \frac{pc}{s} ) \cdot (1 \times 10^{-6} \frac{Mpc}{pc} ) ,\] which yields: \[H_0 = 2.27 \times 10^{-18} \frac{1}{s} ,\]  and once this is plugged into the equation for the age of the universe we found a just a bit ago, we now have: \[t = \frac{1}{2.27 \times 10^{-18}  s^{-1}} ,\] \[ t = 4.41 \times 10^{17} s\] and now we just convert this value into years so it looks like it makes sense: \[ t =  4.41 \times 10^{17} s \cdot (3600 \frac{s}{hour}) \cdot (24 \frac{hours}{day}) \cdot (365 \frac{days}{year}) ,\] \[t = 1.4 \times 10^{10} years ,\] a close approximation to the current definition for the age of the universe, which is 200 million years older than the oldest globular cluster we have observed.

(b) To find the Total Distance of the galaxy, it is as simple as multylping the age of the universe times the maximum speed of the universe, the speed of light. So by taking the speed of light and turning it into the distance it travels in one year: \[ c = 3 \times 10^8 \frac{m}{s} \cdot (3600 \frac{s}{hour}) \ cdot (24 \frac{hours}{day}) \cdot (365 \frac{days}{year}) , \] \[c = 9.46 \times 10^{15} \frac{m}{year} , \] and next we turn the speed of light into Mpc per year: \[ c = 9.46 \times 10^{15} \frac{m}{year} \cdot \frac{1}{3.241 \times 10^23 \frac{m}{Mpc} } , \] \[c = 2.77 \times 10^{-8} \frac{Mpc}{year} , \] and now finally multiply this by the age of the universe and we get: \[1.4 \times 10^{years} ~years ~\cdot 2.77 \times 10^{-8} \frac{Mpc}{year} = 3.87 \times 10^2 Mpc .\] However, this is not the total observable distance of the universe, rather it is only the radius of our line of sight, so to complete the distance, we would need to multiply this radius by 2 to get the total diameter of the sphere that we consider the observable universe, the Hubble Length. \[ Hubble ~Length= 2 \times  3.87 \times 10^2 Mpc ,\] \[Hubble ~Length = 7.7 \times 10^2 Mpc . \]


Introduction to growth in infinite directions

We are bounded in a nutshell of Infinite space: Blog Post #26, Worksheet # 8.1, Problem #1: Introduction to growth in infinite directions

1. Before we dive into the Hubble Flow, let’s do a thought experiment. Pretend that there is an infinitely long series of balls sitting in a row. Imagine that during a time interval \(\Delta t\) the space between each ball increases by \(\Delta x\).


(a) Look at the shaded ball, Ball C, in the figure above. Imagine that Ball C is sitting still (so we are in the reference frame of Ball C). What is the distance to Ball D after time \(\Delta t\)? What about Ball B?

(b) What are the distances from Ball C to Ball A and Ball E?

(c) Write a general expression for the distance to a ball N balls away from Ball C after time \(\Delta t\). Interpret your finding.

(d) Write the velocity of a ball N balls away from Ball C during \(\Delta t\). Interpret your finding.

(a) As Edwin Hubble trained his telescope onto the greater heavens, objects thousands if not hundreds of thousands of light years away, he began to see the hints of something in play in the universe, something few had ever thought of conceiving as possible. Since ancient times, the Earth was the center of the cosmos, with everything rotating around it, an idea disproven during the Scientific Revolution. But even these great minds did not envision how humanity would later understand how there were galaxies and clusters and unimaginably large structures in the cosmos which left Hubble with a clear idea: the universe was growing, every second, of every day. The expansion Hubble saw can be described by various methods, mainly by his own Hubble Constant, but more on that in the next post.

First let us discuss how this expansion works, how objects that are spaced from each other all move at a same speed, but because of their original place, they continually look farther and farther away.

In the example of a row of balls, we have the question of how much distance separates B and D from C, in its point of view. For each case, the distance to B or D is \(\Delta x\), as is established by the question after a time \(\Delta t\).

(b) The same reasoning applies to balls A and E, where if the space from B to A has to be \(\Delta x\) and if the distance from C to B is already \(\Delta x\), then the distance from C to A has to be \(2 \Delta x\), a case identical to ball E.

(c) So, if the balls’ position away from the point of reference is the guiding force driving the distance it has after a time \(\Delta t\), a ball N balls away will always be \(N \Delta x\) away after a time \(\Delta t\). So the equation for the distance from the ball C will always be: \[D(\Delta t) = N\Delta x .\] This shows, and we’ll get more into this in just a bit, that the farther an object was at the start, the farther it will move after a time \(\Delta t\)

(d) Now that we have a general equation for the distance from the point of reference, we can use it to find the velocity of the object for any original condition. From kinematics we already know the equation for distance traveled by an object with constant velocity is: \[X_f = X_i + v(\Delta t) ,\] which can be rewritten as: \[ v = \frac{X_f – X_i }{\Delta t} ,\] \[ v = \frac{\Delta x}{\Delta t}, \] which happens to be very similar to the equation for distance we have already found. Now we simply plug in the actual distance equation for the case of a constant expansion as we have been describing: \[v = \frac{N \Delta x}{\Delta t}, \] which describes the velocity of an object we are viewing from a particular reference point. From it, we can clearly see that an object that is farther away is perceived to be moving increasingly faster, a fact which has helped astronomers many times.

As we will see in the next post, apparent and actual velocities help a great deal, permitting us to understand just how far away an object is (giving us a new rung on the distance ladder to use as necessary), as well as how long its light has been traveling, but more on this next time.