Está en la página 1de 16

Shang Wang

od-asteroid6249jennifer.weebly.com

Orbital Determination of the Asteroid 6249 Jennifer


Abstract
Asteroids can cause major destruction when they collide with a planet. My team’s goal was to
calculate the Mars-crossing asteroid 6249 Jennifer's orbit in order to prevent possible collisions with
Mars, a candidate for human colonization.
We prepared several observations each night using TheSkyX software, the SAOImage DS9
application, and the JPL Horizons database. After imaging 6249 Jennifer at the Sommers-Bausch
Observatory, we processed the light frames to remove bias and used Least Squares Plate Reduction to
determine its right ascension and declination on 18 different dates. We then utilized Gauss’s Method
to compute the orbital elements of our asteroid. We also used several other methods to ensure the
consistency and accuracy of our results.
However, due to the minor gravitational tug of 6249 Jennifer with other objects in the Solar
System, Gauss’s method alone does not show the full picture. We used a simulation program, Swift,
to perform long-term integrations. Results show that the asteroid 6249 Jennifer has a 72.7%
likelihood of staying alive after 40 million years, a relatively stable orbit. It is necessary to continue
the comprehensive, ongoing astronomical survey so that if 6249 Jennifer is on a collision course with
Mars, we will have enough time to deflect the asteroid away.

Team contributions
In a team of three, we operated the telescope and collected data together at the
Sommers-Bausch Observatory. The original project idea was assigned by the Summer Science
Program. However, when choosing our asteroid, I suggested to my teammates that we should
specifically research the asteroid 6249 Jennifer for the following reasons: it had limited information
in celestial databases relative to other possible asteroids to study; and throughout the 6 weeks of
research, 6249 Jennifer transited between 11:00 pm and 1:30 am, giving us the perfect viewing
opportunity. My idea was accepted. My teammates and I later found that based on NASA’s JPL
Horizons ephemeris, 6249 Jennifer had a high declination, which would allow us to better view the
asteroid through a thinner layer of atmosphere.
Since I had prior experience in utilizing various Python libraries, I wrote a major portion of
our data-analysis software in Python that was used to calculate 6249 Jennifer’s orbit. Specifically, I
coded Gauss’s Method, which was the most crucial part of our project aside from observation. I also
took lead in my group’s weekly research meetings with our research supervisors, presenting our
progress and creating our weekly work agenda. Yet, this project was such a complicated process that
we could not hope to tackle it alone; the three of us approached the project from different angles and
directions in accordance to our individual skill sets and strength.

What I learned
I’ve always been fascinated by the endless possibilities of our universe. I’m thankful that SSP
introduced me to hands-on astrophysics research. I learned how to be more methodical, patient, and
mindful during the scientific process.
Also, during our first two observations, we kept getting null results. We were still confused
even after long hours of literature searches. Finally, in our 3rd observation, we realized that we had
forgotten to set the equinox to the J2000 epoch, a different reference point in time. Who knew that
even one small mistake would lead to null results in research? In challenges like these, my teammates
and I bonded. I will never forget the sense of accomplishment we felt when our code ran or when our
star charts matched.

-1-
Shang Wang
od-asteroid6249jennifer.weebly.com

1. Introduction
Our solar system is characterized by a complexity that we are only beginning to understand the
true magnitude of. Consisting of at least eight planets, a plethora of dwarf planets, and an
uncountable number of minor objects, the countless number of objects comprising our solar
system — and the interactions between them — are subject to continual research and observation
as we continue to refine our understanding.

Of these objects in the Solar System, the least understood are the minor planets: objects not large
enough to warrant the assignment of dwarf planetary or planetary status, but still large enough
for individual objects to be detected from Earth. Objects of this category can range from a few
meters to hundreds of kilometers across and are typically partitioned into two distinct categories:
comets and asteroids. Comets are objects primarily comprised of water ice, frozen gases, and
dust. Asteroids, by contrast, are solid bodies that typically possess carbon-rich, metallic
compositions. There is no concrete definition for an asteroid other than a sun-orbiting object that
is neither a recognized planet nor a comet; as a result, objects recognized as minor planets and
asteroids frequently overlap (Ceres is both an IAU-recognized dwarf planet and an asteroid, for
example).

Asteroids and comets can both be found in virtually every part of the solar system; however,
comets are generally found in the outer solar system (beyond the orbit of Neptune). Asteroids
can be found in highest concentrations in the asteroid belt, a region of space between the
trajectories of Mars and Jupiter. However, they can also be found orbiting in large quantities at
the Lagrange points of Jupiter (Trojans), orbiting within the orbits of the giant planets
(Centaurs), and orbiting within the Kuiper belt and other regions well beyond the orbit of
Neptune.

Of particular interest to us are asteroids that orbit within the inner solar system, inside the inner
boundary of the asteroid belt. These are asteroids that often intersect the trajectories of the rocky
planets; of these, we define Earth-crossers, asteroids whose orbits cross the semi-major axis of
the orbit of Earth, and Mars-crossers, asteroids whose orbits cross the semi-major axis of Mars.
Earth-crossers carry heightened importance due to their potential to impact the Earth (provided
their orbits sufficiently intersect), with the largest asteroids posing a substantial threat to life on
Earth and human civilization, and medium-sized asteroids (those with radii of 20-100 meters)
devastating massive swaths of land. The majority of asteroids that have impacted the Earth are
very small, but rarely, large ones impacted the Earth as well and threatened the existence of an
entire species. The Earth has had approximately five mass extinctions in the course of its
existence, one of which is the Cretaceous-Paleogene extinction event with a hypothesis that
suggests the cause to be a large asteroid [1]. Yet, disaster prevention is only part of our
motivations for tracking asteroids. Asteroids are also believed to be the source of how water
came to Earth and consequently, the source that brought Earth life [2].

Therefore, asteroids are subject to considerable tracking and monitoring efforts by space
agencies across the globe. For example, European Space Agency built the space probe Rosetta

-2-
Shang Wang
od-asteroid6249jennifer.weebly.com

which flew by two asteroids (2867 Steins and 21 Lutetia) and helped scientists gain a better
understanding of events that happened millions of years ago [3]. Notable examples of other
efforts include the Catalina Sky Survey, a comprehensive, ongoing astronomical survey which
has discovered 3 small asteroids just prior to impacting Earth [4]; and the Near Earth Asteroid
Tracking (NEAT) program, a NASA-JPL collaborative between observatories in Hawaii and
California [5].

The specific target of our research is the asteroid known by its provisional designation 6249
Jennifer, a Mars-Crossing asteroid. While Mars-Crossing asteroids do not generally elicit the
same levels of saliency that Earth-Crossing asteroids do, they are still of significant interest to
space agencies due to Mars's status as a promising candidate for potential future human
colonization. It is necessary to continue the comprehensive, ongoing astronomical survey so that
if 6249 Jennifer is on a collision course with Mars someday, we will have enough time to deflect
the asteroid away.

Fig. 1. Model of 6249 Jennifer orbiting in 3D space, with its trajectory shown relative to those of
the Sun and the four planets closest to the Sun. It is visible that the orbit of 6249 Jennifer crosses
the orbit of Mars [8].

The primary goal of our research is determining the orbit of 6249 Jennifer, which hinges around
the determination of its six orbital elements: the semi-major axis (a), the average distance from
the asteroid to the sun; the orbital eccentricity (e), the elongation of the ellipse of the asteroid's
orbit; the inclination (i), the angle between the plane of the asteroid's orbit and the sun's
equatorial plane; the longitude of the ascending node (Ω), the position along the asteroid's orbit
at which the asteroid's orbital plane intersects the ecliptic plane; the argument of perihelion (ω),
the angle between the perihelion and the ascending node; and finally the mean anomaly (M), the
angular displacement from the perihelion if the asteroid was following a circular orbit about the
center of its orbital ellipse. See Fig. 2 for a visual of the orbital elements.

-3-
Shang Wang
od-asteroid6249jennifer.weebly.com

Fig. 2. Diagram of all 6 of the Keplerian orbital


elements, from Johannes Kepler and his laws of
planetary motion. In this diagram, the orbital plane is
represented in yellow, and the ecliptic plane (or the
reference plane) is represented in gray. [6]

In order to calculate these orbital elements, we used


the method of Gauss [7]. The method of Gauss is an
iterative numerical method that involves using Taylor
series approximations of the asteroid's observed
motion across the sky in order to converge upon
values for its position and velocity vectors, namely r
and ṙ (with the Sun as the origin, in Newton’s
derivative notation), which in turn undergo a series of algebraic operations in order to find
numerical values for the six orbital elements. In order to use the method of Gauss, at least three
observations are required so that the derivative of the asteroid's position (its velocity) can be
calculated. These observations must be roughly evenly spaced apart in order to avoid the
magnification of inaccurate data that may result from small differentials between data sets.

2. Observations and Image Processing


2.1 Methods
Our data were collected at Sommers-Bausch Observatory at the University of Colorado Boulder
(40.0069°N, 105.2728°W, 1656.6 meters above sea level, UTC-06:00 time zone). The telescope
used to image the asteroid was Apollo, a Planewave 20 CDK reflector telescope, along with an
SBIG Astronomical Instruments CCD (charge-coupled device) camera with a Kodak KAF-8300
optical sensor. See Table 1 for our journal of observations.

Table 1. Journal of Observations

Date Time Number of Quality of Images


(UTC) (UTC) Images Taken

June 28th, 05:00- 15 lights, Cloudy, and telescope settings were off due to failure to set equinox to
2018 06:30 10 darks J2000 instead of apparent.

June 30th, 05:00- 15 lights, Good quality. Relatively clear skies. We matched our star charts easily
2018 06:30 10 darks

July 2nd, 05:00- 15 lights, Good quality images, clear skies. We matched our star charts easily.
2018 06:30 15 darks

July 4th, N/A N/A Canceled due to weather


2018

-4-
Shang Wang
od-asteroid6249jennifer.weebly.com

July 6th, N/A N/A Canceled due to weather


2018

July 8th, 05:00- 15 lights, Mostly clear weather, a few clouds, we matched our star charts easily.
2018 06:30 15 darks

July 10th, 04:45- 15 lights, Clear weather. Although our star chart was generated with a time that
2018 06:00 15 darks was about 20 minutes off from ideal, we were still able to match our star
charts.

July 12th, 04:45- 15 lights, Only a bit cloudy at first, but more clouds began rolling in of the way
2018 06:00 15 darks through our shift. We matched our star charts easily.

July 14th, 04:45- 10 lights, Very cloudy and hazy at first. Could not focus on a star until 10 min
2018 06:00 10 darks have passed. After 37 min, clouds thinned out. We finally took clear
images and matched our star charts easily.

July 16th, N/A N/A Canceled due to weather


2018

July 18th, 04:45- 15 lights, Weather looked 90% clear. A few clouds were on the horizon. We
2018 06:00 15 darks matched our star charts easily.

July 20th, 04:45- 10 lights, Clear weather. At first, we made the mistake of forgetting to move the
2018 06:00 10 darks flip mirror into the CCD position, but we immediately fixed this
mistake. We matched our star chart easily.

July 22th, N/A N/A Canceled due to weather


2018

Before observing each night, we used the JPL Horizons web interface to obtain a predicted right
ascension (RA) and declination (DEC) to help locate 6249 Jennifer [8]. From these predicted
coordinates, we used the astronomical imaging database DS9 to generate a star chart centered on
the predicted location. When observing, we slewed the telescope to the predicted right ascension
and declination of our asteroid, making sure to match our star charts from DS9 with the images
taken from the CCD camera on the telescope.

After observing, several steps were taken to process the raw images in order to obtain data that
could then be used for the orbit determination. The first step was to calibrate the raw light
frames. For this, we used the astronomical imaging software MaximDL. In MaximDL, the
appropriate dark frames, flat frames, and bias frame were loaded into the program. Then, the raw
light frames were calibrated using the darks, flats, and bias frames, and the 15 newly processed
images were animated so that our asteroid was easily visible moving across the screen. In order
to move on to the next stage in image processing, we selected 3 images of out of the 15 from
each observing session to be further analyzed. See Fig. 3. for an example of a calibrated image of
6249 Jennifer that was selected at this step.

The next step after calibrating our raw images was to complete the processes of astrometry and
photometry. Using DS9, we generated star charts for the given night of observations and used the

-5-
Shang Wang
od-asteroid6249jennifer.weebly.com

USNO UCAC4 optical catalog to select a set of 6-8 reference stars with known right ascensions
and declinations. Not all of our viable reference stars had known visual magnitudes; however, we
made sure to have enough reference stars with known visual magnitudes such that the process of
photometry could be carried out. With our reference stars selected, we used MaximDL to
determine the x and y coordinates of each star's centroid. After calibrating the image using one of
the reference star's known visual magnitudes, we located our asteroid in the image and recorded
the x and y coordinates of its centroid as well as its visual magnitude. This process of
determining x and y coordinates of each reference star and our asteroid, as well as determining
the magnitude of our asteroid, was repeated for each of the 3 images for each observation night.
The entire process — starting with generating a star chart in DS9 then selecting the reference
stars through determining the coordinates and magnitude of our asteroid — was repeated for all 6
nights of observations.

Finally, to complete the process of astrometry, we used the method of Least Square Plate
Reduction to generate the right ascension and declination of our asteroid in each image from the
right ascensions, declinations, x-coordinates, and y-coordinates of the reference stars in each
image. The statistical method essentially squares the distance between the calculated residual and
the line of best fit and minimizes those residuals squared.

Fig. 3. A calibrated image of 6249 Jennifer on July 10th, 2018 at 05:13:22.770 UTC

-6-
Shang Wang
od-asteroid6249jennifer.weebly.com

2.2 Results

We list the right ascension (RA), declination (DEC), and apparent magnitude (Vmag) of the
asteroid in Table 2.

Table 2. RA, DEC, and Vmag of asteroid observations with time both in UTC and Julian date.
Note that the local time of these observations was UTC-6:00.

UTC Date Julian Date RA (hh mm ss) DEC (dd mm ss) Vmag

2018-07-02 05:25:26.380 2458301.726 18 16 29.26 24 00 08.7 15.8

2018-07-02 05:46:08.379 2458301.741 18 16 28.17 24 00 17.0 15.6

2018-07-02 06:05:51.389 2458301.754 18 16 27.13 24 00 25.0 15.6

2018-07-08 05:13:20.130 2458307.718 18 09 18.21 24 46 49.3 15.5

2018-07-08 05:30:13.089 2458307.730 18 09 17.35 24 46 53.5 15.5

2018-07-08 05:47:52.049 2458307.742 18 09 16.46 24 46 57.9 15.4

2018-07-10 04:56:59.150 2458309.706 18 07 00.61 24 57 23.4 16

2018-07-10 05:13:22.770 2458309.718 18 06 59.80 24 57 26.7 15.9

2018-07-10 05:29:00.740 2458309.728 18 06 59.03 24 57 29.8 15.8

2018-07-12 05:04:54.810 2458311.711 18 04 45.93 25 05 36.2 16.1

2018-07-12 05:20:44.670 2458311.722 18 04 45.17 25 05 38.6 16.2

2018-07-12 05:39:51.559 2458311.735 18 04 44.26 25 05 41.4 16.2

2018-07-18 04:56:04.679 2458317.706 17 58 35.94 25 15 58.4 15.2

2018-07-18 05:12:34.560 2458317.717 17 58 35.25 25 15 58.5 15.3

2018-07-18 05:30:07.660 2458317.729 17 58 34.52 25 15 58.5 16.4

2018-07-20 05:02:03.539 2458319.710 17 56 45.34 25 14 57.4 16.3

2018-07-20 05:18:17.210 2458319.722 17 56 44.70 25 14 56.7 16.4

2018-07-20 05:34:38.289 2458319.733 17 56 44.06 25 14 55.9 17

-7-
Shang Wang
od-asteroid6249jennifer.weebly.com

3. Orbit Determination
3.1 Methods
The entirety of the orbital determination data analysis software was written in Python. The
algorithm we constructed followed the method of Gauss [7], which was determined in advance to
be compatible with our asteroid's orbit in that its spherical position vectors converge to a finite
numerical solution.

The Python program we wrote accepted inputs of three separate sets of right ascensions (RA),
declinations (DEC), and Julian times corresponding to given sets of observation data, in addition
to sun position vectors (relative to Earth) at the same points in time as our referenced observation
dates. Given those values, our software returned the six orbital elements of the object in question
(in this case, 6249 Jennifer, although our scripts can certainly generate orbital elements for other
objects as well). For the determination of our asteroid's orbital elements, we used observation
data collected at three unique times on July 2nd (at 5:25 UTC, 5:46 UTC, and 6:05 UTC), three
unique times on July 10th (at 4:56 UTC, 5:13 UTC, and 5:29 UTC), and three unique times on
July 18th (at 4:56 UTC, 5:12 UTC, and 5:30 UTC). We define our “1st dataset” as the set of the
first observations on each of the three dates. Similarly, the “2nd dataset” is the set of the second
observations on each of the three dates, and the “3rd dataset” is the set of the third observations
on each of the three dates. See Table 3 above for a visual of the chosen datasets. The Gaussian
method works best with evenly spaced observations over the course of at least two weeks, so we
used images from these three dates that have most uniform separation out of all six observations.
We calculated through our Python program the orbital elements of our asteroid three times. The
first time we used the data from the first observation on all three dates, the second time we used
data from the second observation on all three dates, and the third time we used data from the
third observation on all three dates. This yielded three lists of orbital elements, shown in Table 9
of the Appendix. To get our final orbital elements list, we took the mean of those three lists
reported in Table 5 below. The only element that we could not compute an average for (as it is
constantly changing depending on time) was M, the mean anomaly of the asteroid. Thus, we
simply reported our three calculations for M. However, in order to assure that our mean anomaly
values were accurate, from each of the three values for M, we calculated T, the time of perihelion
passage, which is a constant characteristic of our asteroid's orbit. Thus, by calculating the mean
value and standard deviation of T, reported in Table 5, we were able to confirm how accurate our
computed values of M were.

Input July 2nd July 10th July 18th

1st dataset 5:25 UTC 4:56 UTC 4:56 UTC

2nd dataset 5:46 UTC 5:13 UTC 5:12 UTC

3rd dataset 6:05 UTC 5:29 UTC 5:30 UTC

Table 3. The three datasets chosen to be the input for the software.

-8-
Shang Wang
od-asteroid6249jennifer.weebly.com

In determining the accuracy of our Python programs, we performed a self-consistency test. For
this test, we took the right ascension and declination for one of our observations, July 20th at
5:34:38.289 UTC, as our test value. Next, we took the list of mean orbital elements as seen in
Table 5 and ran it through an ephemeris generator that we had programmed in Python. We also
gave the ephemeris generator the Julian date and Earth-Sun Vector values for that observation
time, which returned the predicted right ascension and declination for the test case observation
time, as expected. Table 4 shows the original and predicted right ascension and declination
values for that observation, along with the percent error between them. From the very small
percent differences viewed in this situation (on the magnitude of 10-5 and 10-2), we can confirm
that our program accurately returns a predicted right ascension and declination for a given date
based on the list of orbital elements.

Table 4. Results of self-consistency test


Position Original Value Processed Value Percent Difference

Right Ascension (hh mm ss) 17 56 44.06 17 56 44.03 3.776 x 10-5

Declination (dd mm ss) 25 14 55.9 25 14 54.6 1.460 x 10-2

3.2 Results

The mean value of 5 of the 6 orbital elements of 6249 Jennifer is in Table 5, along with the mean
value of the time of perihelion passage (T), with estimates of uncertainties in the standard format.
*The only element that we could not compute the mean value for was M, the mean anomaly of
the asteroid because it is constantly changing. Thus, we simply reported our three calculations
for M. For the table with the six orbital elements of 6249 Jennifer from our three data sets, along
with the time of perihelion passage before taking the mean, see Table 8 in the Appendix section.

Table 5. Mean values of orbital elements of 6249 Jennifer. Note that the Julian date listed to the
right of each M is that of the middle observation in the set of data used to calculate that M.
Orbital Element Mean Value with Uncertainty Standard Deviation

a (AU) 1.91424±0.00024 0.00041467

e 0.14188±0.00041 0.00070758

i (degrees) 28.09607±0.00516 0.0089386

Ω (degrees) 221.67866±0.02029 0.035151

ω (degrees) 151.57538±0.01811 0.031375

M (degrees) for JD 2458309.706 287.34245 Not meaningful *

M (degrees) for JD 2458309.718 287.49273 Not meaningful *

M (degrees) for JD 2458309.728 287.36771 Not meaningful *

-9-
Shang Wang
od-asteroid6249jennifer.weebly.com

T (Julian days) 2457537.431±0.26723 0.46286

4. Discussion of Orbit Determination


After running our orbital determination code on our asteroid, the mean orbital element values we
obtained were nearly identical to those listed in JPL’s Horizons database. Our values are
compared to those from JPL in Table 6 below. For a comparison of our calculated orbital
elements for all three sets of data with those of JPL, see Table 9 in the Appendix section.

Table 6. Our calculated mean orbital elements of 6249 Jennifer compared to those from JPL
Horizons. [8]

Orbital Element Calculated Value JPL Value Percent Error

a (AU) 1.91424 1.914739 0.026125%

e 0.14188 0.141743 0.099531%

i (degrees) 28.09607 28.108278 0.043435%

Ω (degrees) 221.67866 221.693815 0.0068368%

ω (degrees) 151.57538 151.575183 0.00012768%

T (Julian days) 2457537.431 2458504.906 0.039360%

When performing the self-consistency check, we were able to recover our observed right
ascension and declination with a percent difference on the magnitude of 10-5 for right ascension
and 10-2 for declination. With such a small percentage difference, we are confident that when
given a list of orbital elements, our program accurately predicts the right ascension and
declination for our asteroid on a given date.

The quality of our results may have been influenced by factors beyond our control. For instance,
atmospheric turbulence may have slightly shifted the location of either our asteroid or the
reference stars we used to calculate our asteroid's position. If such shifts had taken place, the
final right ascension and declination of our asteroid would not be completely accurate, as the
data we had relied on in the process of calculating its right ascension and declination would have
been slightly skewed. A fix for this issue would be to perform the experiment outside of earth's
atmosphere, where atmospheric turbulence could not shift the apparent locations of astronomical
objects. However, that change would require a significant increase in funding as it would require
launching equipment into orbit around the earth.

Another factor which may have altered our results was the fact that we used a master bias frame
to calibrate our light frames, rather than using a bias frame specific to each individual light
frame. The master bias frame is the average of several bias frames, thus it is a good
representation of what the bias of an image likely is. However, the actual bias of an image varies

- 10 -
Shang Wang
od-asteroid6249jennifer.weebly.com

from the light frame to light frame, so calibrating a raw light image using the master bias may
not accurately account for the actual bias. The result of this is that some calibrated light frames
may not have the exact right counts. Therefore, the magnitude of our asteroid and the reference
stars may have been slightly skewed. In order to fix this issue, we could have taken a bias frame
after every light frame we took. This still would not have been as accurate as instantaneously
taking a bias frame while the light frame was being taken, but it would be more accurate than
just using a master bias frame for all of the images.

5. Orbit Integration
We proceeded the Orbital Determination Project with a follow-up project. The goal was to
calculate the long-term dynamical behavior of 6249 Jennifer once we have determined its current
osculating Keplerian Orbital Elements.

Kepler's 1st Laws of Planetary Motion states that objects follow an elliptical orbit about the Sun,
where the trajectory can be characterized by six orbital elements (See Fig. 2 for a diagram): the
semi-major axis (a), the orbital eccentricity (e), the inclination (i), the longitude of the ascending
node (Ω), the argument of perihelion (ω), and the mean anomaly (M). We used the Method of
Gauss — a two-body computation, which is only accurate if there are two bodies in the system.
Fortunately, this approximation works effectively because the Sun is 1000 more massive than
any other objects in the Solar System.

However, this two-body approximation fails over long timescales of if the object encounters
close approaches with another object much more massive. For example, Fig. 4 below shows the
evolution of the Earth’s orbital inclination and eccentricity in a simulation of Earth over the next
600,000 years [9]. These values undergo slow (relative to Earth’s orbital period) oscillations due
to the gravitational tugs of other planets. These changes are the cause of climate variations we
see on Earth. The effect of the planets on minor objects, particularly those that cross the
trajectories of planets in the Solar System (such as 6249 Jennifer, a Mars-crossing asteroid), can
be even more dramatic. During the next millions of years, major masses such as the Sun,
Mercury, Venus, the Earth-Moon system, Mars, Jupiter, Saturn, Uranus, and Neptune may skew
the orbit of 6249 Jennifer to such an extent that may lead to the eventual demise of the asteroid.
Our goal is to determine how the orbit of 6249 Jennifer changes over the next tens of millions of
years, perhaps also predict the eventual fate of our asteroid.

Fig. 4. Simulation of Earth’s inclination (i) and eccentricity (e) over the next 600,000 years. [9]

- 11 -
Shang Wang
od-asteroid6249jennifer.weebly.com

5.1 Methods
We used Swift, a solar system integration software with packages of different complex
algorithms, to perform long-term integrations on a set of mutually gravitationally interacting
bodies [10]. Swift is a version of the Regularized Mixed Variable Symplectic (RMVS) method
[10]. We focused on 6249 Jennifer’s interactions with the Sun, Mercury, Venus, the Earth-Moon
system, Mars, Jupiter, Saturn, Uranus, and Neptune since they may be the main causes of
perturbations to its trajectory.
We created three input files for 18 test asteroids/particles (Note that we will be using the terms
“test asteroids,” “test particles,” and “small bodies” interchangeably). These files inform Swift
procedures to perform the calculations (time duration, method of outputting the calculated
results, etc.), initial conditions of the planets (original positions and velocities), initial conditions
of the small bodies (original positions and velocities), as well as other variables that keep track
of the test asteroids. In order to obtain the positions (x, y, z) and velocities (vx, vy, vz), we
generated the ephemeris using the JPL Horizons Web-Interface [8].
We generated 18 test particles using the procedures and varied the magnitudes and directions of
the offset in the input files so that all asteroids are different. The offsets represent the different
gravitational perturbation.
5.2 Results
We generated graphs of aphelion distance versus time (Blue), semi-major axis length versus time
(Black), and perihelion distance versus time (Magenta). Most of the resulting graph shows the
test particles being relatively stable. See Fig. 5-8 for visual.
Test Particle #1 is one of the few test particles that are unstable. In Fig. 9, Test Particle #1 shows
close interaction with Mars at 4.5 million years. The black line representing the semi-major axis
crosses 1.524 AU, which is the semi-major axis of Mars. The interaction caused skewness to the
particle’s orbit since the graph seems chaotic after 5 million years.
The test particles obtained some interesting cases as well. According to Fig. 10, Test Particle
#8’s orbit become circular since its aphelion decreased and perihelion increased. It is interacting
with some unknown outside bodies and increasing angular momentum. Since its angular
momentum increased, its velocity has likely increased as well.
Fig. 5-6 Graphs for Test Particles #2 and #4

- 12 -
Shang Wang
od-asteroid6249jennifer.weebly.com

Fig. 7-8 Graphs for Test Particles #5 and #7

Fig. 9, 10. Graphs for Test Particles #1 and #8

Our results show that the asteroid 6249 Jennifer has a relatively stable orbit and has a 72.7%
likelihood of staying alive after 40 million years. 13 out of the initial 18 test particles remained
alive, and of the 5 particle deaths, 2 were caused by being too far from the sun and 3 were caused
by having too small of a perihelion. See Fig. 11 below for pie chart of particles’ fates. See Table
7 for more information on particles’ deaths.

Fig. 11 Relative Frequency

- 13 -
Shang Wang
od-asteroid6249jennifer.weebly.com

Table 7. Information of particles that came to their eventual demise.

Particle Cause of demise Last planet Semi-major Perihelion Aphelion Inclination


encountered axis (AU) (AU) (AU) (AU)

#1 Perihelion too small Mars 2.5597 0.0029 5.1165 21.8935

#8 Position too far Jupiter 3282.9741 4.8789 6561.0693 38.9966


from Sun

#11 Perihelion too small Mars 2.5597 0.0029 5.1165 21.8935

#12 Perihelion too small Earth-Moon 2.0568 0.0046 4.1091 45.9108


system

#14 Position too far Jupiter 562.1436 9.4103 1114.8769 26.5106


from Sun

6. Discussion of Orbit Integration


The Swift simulation supports the classification of 6249 Jennifer as an asteroid in the Hungaria
group, which is a dynamical group of asteroids in the asteroid belt. The Hungaria group consists
of asteroids that have a semi-major axis between 1.78 and 2.00 AU. All asteroids in this group
have stable orbits because the eccentricity of all are below 0.18 AU. We are not surprised that
6249 Jennifer has 72.7% likelihood of survival as its orbit will remain stable for tens of millions
of years into the future.

However, the Swift simulation only accounted for the largest masses in the Solar System that
could have gravitational effects on our test particles, such as the eight planets and the Sun. Thus,
the results may not be accurate since they do not fully portray all future scenarios for 6249
Jennifer. Other objects in space can also affect the life of an asteroid. For example, 6249 Jennifer
could collide with another asteroid of the asteroid belt in its vicinity due to strong gravitational
effects. This would send debris of the collision flying in unpredictable manners. However, this
scenario cannot be simulated as the number of asteroids that will cross paths with 6249 Jennifer
is unknown. Space, even within our solar system, is too vast with too many objects to simulate
all possible cases with this program.

Space, with its uncountable uncertainties is too complex for us to fully predict, however, this
only motivates us to keep track of as many objects as we can and as frequently as we can. In this
particular case for the Mars-crossing asteroid 6249 Jennifer, it is necessary to continue the
comprehensive, ongoing astronomical survey so that if 6249 Jennifer is on a collision course
with Mars someday, we will have enough time to deflect the asteroid away.

- 14 -
Shang Wang
od-asteroid6249jennifer.weebly.com

Acknowledgment of Major Assistance


I would like to acknowledge my teammates Eva Smerekanych and Logan Keifling for their
significant support, collaboration, and dedication on this project support at 2018 Summer
Science Program. I am grateful for my academic directors, Dr. Agnes Kim and Dr. Michael
Dubson, without whom, I would have lacked the knowledge to complete this project. I appreciate
our teaching assistants, Dahlia Baker, Maria Camila Remolina Gutierrez, Bradley Emi, and Isaac
Guerrero, for their continuous support and assistance. Additionally, I would like to thank Fabio
for providing us the workspace of Sommers-Bausch Observatory at the University of Colorado
Boulder.

References
[1] Cowen, R. (2000). THE K-T EXTINCTION, History of Life. Retrieved July 1, 2018,
from http://www.ucmp.berkeley.edu/education/events/cowen1b.html
[2] Choi, C. Q. (2017, September 21). Asteroids: Fun Facts and Information About
Asteroids. Retrieved August 1, 2018, from https://www.space.com/51-asteroids-
formation-discovery-and-exploration.html
[3] E. S. Agency. (2017). Debris of the Solar System: Asteroids. Retrieved July 1, 2010,
from http://www.esa.int/Our_Activities/Space_Science/Rosetta/Debris_of_the_Solar_
System_Asteroids
[4] Lunar & Planetary Laboratory, U. (2018). Catalina Sky Survey. Retrieved 2018, from
https://catalina.lpl.arizona.edu/
[5] Stokes, G. H., Evans, J. B., & Larson, S. M. (2002). Near-Earth Asteroid Search
Programs. Asteroids III, W. F. Bottke Jr., A. Cellino, P. Paolicchi, and R. P. Binzel (eds),
University of Arizona Press, Tucson, 45-54. Retrieved July 24, 2018, from
http://adsabs.harvard.edu/abs/2002aste.book...45S
[6] Keplerian orbital elements. (n.d.). Retrieved July 24, 2018, from http://www.planetary.org
/multimedia/space-images/charts/diagram_orbit_ephemeris_definitions.html
[7] Boulet, D. L. (1991). Methods of orbit determination for the microcomputer. Richmond,
VA: Willmann-Bell.
[8] NASA. (n.d.). JPL Horizons web-interface. Retrieved July 1, 2018, from https://ssd.jpl.
nasa.gov/horizons.cgi.
[9] Vieira, L., Norton, A. A., Dudok de Wit, T., Kretzschmar, M., Schmidt, G., & Cheung,
M. (2012, August 29). How the Inclination of Earth's Orbit Affects Incoming Irradiance.
Geophysical Research Letters, VOL. 39, L16104. Retrieved July 24, 2018, from
https://pubs.giss.nasa.gov/docs/2012/2012_Vieira_vi06000k.pdf.
[10] Kaufmann, D. E. (n.d.). SWIFT, A solar system integration software package. Retrieved
July 26, 2018, from http://www.boulder.swri.edu/swifter/

- 15 -
Shang Wang
od-asteroid6249jennifer.weebly.com

Appendix
Table 8. 6 orbital elements of 6249 Jennifer from our three data sets, along with the time of
perihelion passage.

Orbital Element 1st Dataset 2nd Dataset 3rd Dataset Mean Standard Deviation

a (AU) 1.91404 1.91472 1.91396 1.91424 0.00041

e 0.14139 0.14269 0.14157 0.14188 0.00071

i (degrees) 28.10370 28.08624 28.09829 28.09607 0.00894

Ω (degrees) 221.70454 221.63864 221.69280 221.67866 0.03515

ω (degrees) 151.55063 151.61066 151.56483 151.57538 0.03138

M (degrees) 287.34245 287.49273 287.36771 Not meaningful Not meaningful

T (Julian Day) 2457537.70 2457536.90 2457537.70 2457537.43 0.46300

Table 9. Comparison of our calculated orbital elements for all three sets of data with those of
JPL. Note that the Julian date listed below each dataset name is that of the middle observation in
each set. [4]

Orbital Element 1st Dataset 2nd Dataset 3rd Dataset


(JD 2458309.706) (JD 2458309.718) (JD 2458309.728)

a % difference 0.0242008809551 0.011121001642 0.0282569528985

e % difference 0.261660552657 0.660430035494 0.131905404253

i % difference 0.0164715555518 0.0785993596225 0.0357486532248

Ω % difference 0.00482529177658 0.0248998277055 0.000471877382463

ω % difference 0.0162230671976 0.0233819285521 0.00685655451301

M % difference 0.0105436584346 0.0401985213781 0.00460050484342

T % difference 0.039341317319 0.0393739040411 0.0393412714985

- 16 -

También podría gustarte