This is the second of three very dense lessons on the fundamentals of solar energy. Keep it up! The lessons will lighten up a bit as we transition past these core materials.
The Sun provides us with shortwave band light at the Earth's surface, and the Earth emits longwave band light up toward the sky; but what is the role of the Sky/Atmosphere in the energy balance on Earth, and how does the atmosphere affect the bands of light that we collect and find useful at the Earth's surface? In solar energy conversion systems, engineers and design specialists apply meteorological information, particularly shortwave band irradiation (from the Sun) to inform project development and design, and also to inform daily performance of installed systems. To a lesser extent (for some radiative thermal estimations), longwave band irradiation from the Earth and atmosphere is used, but the main driver for our SECSs is the shortwave band.
Applied solar energy workers need to use the same science as those working in the field of Meteorology. The shortwave band of light can be collected via ground detectors (pyranometers/pyrheliometers [1]) or from satellites (GOES East [2], GOES West [3]) to probe and diagnose the behavior of light interacting with the atmosphere and the surfaces of Earth over time. Satellites in orbit have instruments called radiometers [4] to detect and measure the intensity (brightness) of the light scattered back to the satellite. Satellite images of visible light display the shading from clouds, the cloudiness of the Earth's surface, and features such as smoke emerging from large forest fires or volcanoes spewing ash. The signal is a measurement of electromagnetic radiation, both the intensity of the light from the Sun, and the ability of the particles or surfaces to reflect that light back up to the radiometer. Think about what determines the brightness of the light that reflects (is "back-scattered") up to the satellite.
We will start with the concept of electromagnetic radiation, including some important bands of electromagnetic radiation invisible to the human eye. We will also point out ways that a meteorologist and climatologist collect and interpret satellite or radar imagery derived from electromagnetic radiation.
As a preview to Lesson 3, please review the learning objectives on the next page to get an initial sense of the overarching goals that encompass this stage in your forecasting apprenticeship.
At the end of this lesson, you should be able to:
This lesson has two learning activities and will take us two weeks to complete. Please refer to the Course Calendar in Canvas for specific timeframes and due dates - those can vary from semester to semester. Specific directions for the assignments below can be found further within this lesson.
Required Reading: |
J.R. Brownson, Solar Energy Conversion Systems (SECS), Chapter 3 - Laws of Light (review) C. A. Gueymard (2012) Clear-sky irradiance predictions for solar resource mapping and large-scale applications: Improved validation methodology and detailed performance analysis of 18 broadband radiative models [5] Solar Energy 86, 2145-2169. |
---|---|
Optional Reading: | Bird and Hulstrom (1981): Simplified Clear Sky Model for Direct and Diffuse Insolation on Horizontal Surfaces, Technical Report No. SERI/TR-642-761, Golden, CO: Solar Energy Research Institute Meteorology for Scientists and Engineers, 2nd ed. by Roland B. Stull (Chapter 7). J.R. Brownson, Solar Energy Conversion Systems (SECS), Chapter 8 - Measure and Estimation of the Solar Resource |
HOMEWORK: | Learning Activities
|
YELLOWDIG: |
Discussion Topic 1: Climate Regimes Discussion Topic 2: Atmospheric interactions - clouds, aerosols, molecules |
QUIZ: | Quiz Assignment: Laws of Light and Climate Regimes (see Canvas) - due by the end of the first week |
If you have any questions, please post them to the Lesson 3 General Questions thread in Yellowdig. I will check these forums regularly to respond. Feel free to go through the comments and post your own responses if you are able to help out a classmate.
While reading, pay attention to the basic rules of light measurement. Our historical affinity for burning fuels has been paired with an affinity and strong awareness of temperature, but we lack an awareness for measures of light, found through radiometry. Think of how you “know” what 80 °F versus 65 °F means. Why not measure the irradiance/irradiation along with the temperature? What if we, as a solar culture, were to learn what 1200 W/m2 meant versus 500 W/m2 for an average hourly or minute irradiation upon an exposed surface (vertical wall or sloped rooftop)?
For solar resource assessment, radiometry (the measure of electromagnetic or radiant energy) is more valuable than thermometry (the measure of temperature). Solar energy is comprised of the shortwave band of light found between 250-2500 nanometer wavelengths of light ( , or one billionth of a meter). Within the solar field, we measure the shortwave band in terms of irradiance, as W/m2 (Watts per square meter, a flux of light per receiving area). If we group irradiance over a block of time, say an hour, we call the measure irradiation, in units of Wh/m2 (Watt-hours per area, similar units to electrical energy measure).
To convert energy values as electron volts from nanometer wavelengths: divide 1239.8 nm•eV by the wavelengths that you found in nanometers to find their respective energies in electron volts (eV). Because the error is negligible for us in the field, we can even use 1234.5 nm•eV (just count to five and use the right decimal point).
For more information on how we get to this equation, see the Key Equations [7] page in the Resources box.
Traditional silicon photovoltaics collect energy for electricity production from wavelengths <1100 nm. Other systems like solar hot water panels will collect almost all of the spectrum from the Sun. In fact, almost all of the opaque surfaces about our environment (buildings, roads, water) tend to absorb strongly over the entire shortwave range (leading to an increase in temperature!).
And yet, our wonderfully adaptive human eyes only capture the tiny visible band from 380-780 nm. Thanks to the complex feedback systems from our irises, lenses, and eyelids, eyes can adapt well to extreme conditions of bright or dim lighting but cannot be used to quantitatively gauge irradiance conditions that energetically drive our buildings and photovoltaic systems. Our eyes impart important information, but because they are highly sensitive and adaptive systems, our eyes cannot be used to reliably distinguish and evaluate the solar resource (bare skin may actually be a better receptor for the shortwave band on days that are not windy).
Notice that the "visibility" of a photon is really limited only by the type of detector being used. Let's review the distribution of photons across the electromagnetic spectrum, and then we will discuss types of photon detectors beyond the rods and cones in our eyes. As a reminder, 1 micrometer wavelengths (also called 1 micron wavelengths) are equivalent to 1000 nanometer wavelengths (by unit conversion). It is often easier to use units of nanometers (nm) for the shortwave band, while using units of micrometers (um) for the longwave band. Either will work, but it is helpful to be aware of the relationship.
Now, let's zoom in to the shortwave and longwave bands important to SECSs. Wavelengths are expressed either in nanometers (nm) or micrometers ($\mu m$, also called microns). For wavelengths greater than 380 nm, electromagnetic radiation transitions from the ultraviolet (UV) to the violet visible range for the human eye. As we will observe in the Granqvist figure below, the red visible range ends for wavelengths greater than 780 nm (to the human eye). Notice how the "infrared" crosses over both the shortwave and longwave bands of light.
All objects glow with energy in proportion to temperature. As temperatures increase, the wavelength oscillations become shorter and the density of photons measured is effectively increased. So, the light becomes more intense (more photons), and each photon is packed with more energy.
Most photons are emitted from a thermal surface, where the constituent atoms are vibrating in the material (this means: temperature). Every material has a thermal property, and hence "everything glows." This concept works well for something called "blackbody emission" (zero emittance) or "greybody emission" (non-zero emittance) but not so well to describe your standard laser pointer or microwave oven. So, just to clarify all cases, there are also photons emitted from the Sun and from solid state devices that are "stimulated emission" (lasers), which can allow for photon emission from surfaces that are high in energy, but not high in thermal temperature.
This is our first principle of seven for the behavior of light. So, let's begin with the simple yet important statement: Light is directional. Light, as a photon, is born (emitted) from a surface in many directions and then impinges upon other surfaces, where it is either absorbed, reflected, or transmitted (and refracted). We use that directionality to describe the resource. In the text chapter, we have defined a visual shorthand for light that is emitted, transmitted, reflected (scattered), and/or absorbed by various surfaces.
Notice that the "Ir-" instructs us that the light is incident upon a surface. For example, we measure "irradiance" on a pyranometer, because light is absorbed by the device. Also, notice that my choice of "-iance" or "-iation" determines if the light is a measure of rate (as in power: energy per time) or energy (rate integrated over a span of time).
All objects glow (even the gases in the sky); it is just the intensity and peak wavelength that shifts with respect to their thermal temperature. This brings us to the second of the seven principles of light behavior. Light (e.g., derived from the sun) has a broad spectrum of energies, which are discussed as wavelengths. We tend to group similar wavelengths in the spectrum according to the properties of their emitting surface or receiving surface.
This is a "group" of high-energy wavelengths of light that are emitted from the Sun (See Light is Directional), comprising the majority of the total energy collected by the Earth's surface or our designed SECS, given the decrease in the light power density with distance to Earth (93 million miles, 150 million km). This bundle of wavelengths of light emitted from the sun includes the ultraviolet (UV), the visible, and the near infrared (IR) sub-bands.
The Sun is effectively the only surface from our surrounding environment that regularly and naturally emits shortwave radiance. We can deliver shortwave band energy to a new surface using a reflector, though, as is the purpose of light concentration onto a central receiver.
Long wave band consists of low energy light. The wavelengths of light that are emitted from our skin, from pavement, the sky, ice cubes, and even hot ovens are coming from surfaces that are quite cool relative to the surface of the Sun, right? We are comparing 250-500K surfaces to a 6000K surface. Planck's Law (discussed in the next page) suggests that the distribution of wavelengths will be very different from cool bodies than from extraterrestrial bodies with internal fusion reactors.
The group of wavelengths from cool bodies on Earth is termed the longwave band. This is >2500nm at Earth's surface, but also can be >3000nm if we were to group the wavelengths from space. The longwave band photons are of lower energy, but there is a very wide band of wavelengths included in the group. The longwave band contributes to the greenhouse effect, and keeps our atmosphere comfortably warm. With the addition of CO2 and water vapor to the atmosphere, we make the sky into a better reflector for keeping longwave energy instead of allowing it to escape into space.
Now, we're going to take a moment to explore the following multiple figure graphic from Smith and Granqvist. There is a lot of information in here, so take your time and revisit it as the class progresses.
Light intensity decreases in proportion to the square of the distance between emitter and receiver. This is called the Inverse Square Law, which applies to solar energy as well as to any other light source. In case of the Sun, the emitted energy flux is 6.33x107 W/m2 at the surface of the Sun drops down to 9126 W/m2 at the surface of Mercury (d=58 million km), and down to 1361 W/m2 at the exterior of Earth's atmosphere (d=150 million km).
As shown in the next figure, the reason for the inverse square law is geometric in nature. As light is emitted from a point, like the Sun at a very large distance, the same quantity of photons is spread out over an increasingly larger area with distance. Area is a spatial term in units of distance squared. This same principle works for ordinary surfaces, as photographers well know.Mathematically, this process of light “dilution” can be represented by the following equation (which is another form of Equation 3.2 in the SECS textbook)
\[G_2^{} = {G_1}\left( {\frac{{d_1^2}}{{d_2^2}}} \right)\]where G1 is the light intensity (or flux) at distance d1, and G2 is the light intensity at distance d2.
If we take G1 as the radiative energy flux at the surface of the Sun, we can then estimate the radiative energy flux at the Earth’s orbit (G2) using this equation simply based on the planetary distances. Try to make this calculation with d1 being the Sun’s radius, and d2 being the distance from the Sun to the Earth and see what you get. You will have a chance to share your result in the Lesson 3 Activity further in this lesson.
We continue our review of the seven basic rules, or "laws," of light. These four are a bit closer to physical laws than what's on the prior page, but we need to understand the basics of light if we are to move on to the physics of light interacting with the atmosphere.
You may want to use the math/science site Wolfram Alpha [12] to solve several of these Self-Test questions quickly, or to check some of the notes in Wolfram Alpha against our own work here. Wolfram Alpha is the Google of math and physics. If you want to do calculations on Wien's Law, you type "wien's law" and you will get a calculator for Wien's Law with all sorts of other information.
This radiative transfer law is very important when considering energy balance. It states that at thermal equilibrium, the emissivity () of a body or surface equals its absorptivity ().
Mathematically, we can conceptualize Kirchoff's law as
The radiant energy emitted from a real surface is represented as E (W/m2), while that of a blackbody (a theoretical condition) is given by EB (W/m2).
Simply put, a surface at steady state temperature will absorb light equally, as well as it emits light. Though light is directional, surfaces exchange photons in both directions.
This law is generalized to mean that all objects have some internal temperature, and given that temperature, they all glow. Max Planck was able to establish the dependence of the spectral emissive energy of a blackbody for all wavelengths of light (), given a known equilibrium temperature of the blackbody.
The Wien's displacement law provides us with expected values for the most probable wavelengths in the Bose-Einstein distribution of blackbody radiation. The law implies that the distribution of photons emitted from a surface at any temperature will have the same form or shape as a distribution of photons emitted from a surface at any other temperature. Mathematically,
The Stefan-Boltzmann Law states that the radiative energy emitted by a surface is proportional to the fourth power of the surface's absolute temperature. The Stefan-Boltzmann Law shows that if you were to integrate all the energies from the wavelengths in Planck's Law, you have an analytical solution of the form below:
where = Stefan-Boltzmann constant
Keep in mind that this is for the surface of the emitter. A surface like the Sun will have a very high value for the energy density on the surface, which then decreases in proportion with the Inverse Square Law from the last page (over the 93 million miles distance to Earth's surface).
Wolfram Alpha website [13] is the Google of math and physics. If you want to do calculations on Wien's Law, you type "wien's law" and you will get a calculator for Wien's Law with all sorts of other information. The purpose of this activity is to practice some basic calculations of physical properties of sunlight.
This video below (6 min) presents a short demo on how to use the Wolfram Alpha site. The Stefan-Boltzmann Law is used as an example here, so it provides you with a tip for solving the second problem.
Prepare a written report with your solutions and upload it to Canvas (Lesson 3 Learning Activity Dropbox) in PDF or docx format. You can use Wolfram Alpha screen shots to present your results for the first two problems (insert the screenshots into your report - don't submit them as separate files). Make sure to present the equations you are using and provide comments to the steps in your solutions.
View the Rubric in Canvas by which this assignment will be graded.
See the Calendar tab in Canvas for specific due dates.
Remember how we stated that light is directional? Light also has a "life" including the birth and death of a photon. Electromagnetic radiation is emitted from a surface and will encounter objects in its many paths. The fate of electromagnetic radiation (as a photon) depends on wavelength of the photon at hand and the physical composition of the objects along its path. Again, the photon is a packet of radiation and can interact with molecules or atoms in its path, being transmitted, absorbed, or reflected.
The action of "reflection" is more appropriately described as "back-scattering" by atmospheric physics, which is why you see that term in the figure of possible paths of light. We also have seen a shorthand diagramming technique from the textbook to describe the life events of ensembles (really big groups) of shortwave and longwave photons.
I would like to pull the concept of photon lifetimes in the following video.
Each of these phenomena is happening simultaneously, and the next video opens up that discussion.
Emission ($\epsilon$), absorption ($\alpha$), reflection ($\rho$), and transmission ($\tau$) can all occur at the same time (in fact it always does). We also know that there is a relationship between light energy balances and temperature. When a receiving surface or material absorbs more energy than it emits, the internal temperature of the material will increase (effectively pumping up the system's energy density). Consider a bright summer day, where your rooftop absorbs much more solar radiation than it emits (or conducts away). The roof surface temperature will rise from just after sunrise to the late afternoon.
Recapping: Electromagnetic radiation (as light) is first emitted, and the emitted photons have an opportunity to be transmitted, absorbed, or reflected by an intervening surface. If that surface is a particle or an assemblage of particles, there will be a tendency for the reflection to be more accurately described as "back-scattering."
In solar energy resource assessment, clouds are the most important atmospheric factor influencing device performance, followed by aerosols (dispersed atmospheric particles).
Case Goal: Develop a working concept of the atmosphere as a case study of a selective surface, interacting in concert with the Earth surface to contribute to the global energy budget. Then, move on to more specific reflective surfaces that we might use as a secondary shortwave resource (diffuse reflectors and specular reflectors) in designing SECS.
The Earth is a vast solar energy conversion system! The atmosphere encasing Earth's land and water mass is a collection of gases and particles that vary in pressure, temperature, and chemistry continuously. If we collectively imagine for just a moment that the atmosphere is a simple cover on top of our main absorber (the Earth), we can begin forming a concept of the atmosphere interacting with electromagnetic radiation across broad bands of wavelengths.
Now, let us consider the optical properties of a single material that reflects light for some bands and transmits or absorbs light for alternate bands, just like the atmosphere represented above. We call the surface of such a material a selective surface, because the light interaction occurs at the surface. A selective surface is non-reflective to some bands of light, while being reflective to other bands of light. The atmosphere is a case study for a selective covering surface between shortwave and longwave bands.
For any given wavelength or band of similar wavelengths, the following three simple phenomena will occur when light interacts with a material surface (where light is either being absorbed or emitted):
This means that each of the simple optical phenomena can be represented by fractions from 0 to 1, with the sum of these equating to 1.
For opaque materials, there is a relation between reflectivity and emissivity (the glow of an object) and between reflectivity and absorptivity.
We now know the relation for surfaces in optics called Kirchoff's Law of Radiation. When a surface is in thermal equilibrium with the surroundings, the emissivity is equal to its absorptivity at each wavelength ($\epsilon = \alpha$). This allows us to make the same relations among reflectivities and absorptivities.
We can show the ways in which the Earth-Atmosphere is selective in the relative properties of each to absorb, reflect, and transmit different bands of light.
We have already described shortwave (280-2500 nm at ground level) and longwave (>2500 nm) bands of irradiation incident upon a surface. The study of optics is that of light-matter interactions, regardless of wavelength. We know that materials like glass are semi-transparent in most of the shortwave band, and materials like pure aluminum are reflective in the shortwave band. We are not so familiar with the way that materials behave in the longwave band, and we often have trouble with materials that are opaque vs. semi-transparent.
Let's review the transmittance graphic that I introduced earlier. This plot has a logarithmic x-axis, so, we can count from 0.2 micrometers (200 nm) up to 1 micrometer wavelengths in increments of 0.1 micrometer. Then, we count from 1 micrometer to 10 micrometers in increments of 1 micrometer, and so on...
We see that from about 0.3 micrometers to about 2.5 micrometers, there is a significant amount of white showing on the plot of the Total Absorption and Scattering, meaning that the absorption and reflection of visible light by the atmosphere is relatively small. In other words, the atmosphere transmits 70-75% of the sun's shortwave light from the top down to the Earth's surface. Along the way, clouds can back-scatter [15] (reflect) some of the visible light into space. In times and places where transmitted sunlight reaches the Earth's surface, then the land, oceans, deserts, grasses, and trees, etc., reflect some of the surface-incident shortwave light back into space (again, with limited absorption along the way).
We also see that from 8-13 micrometers, there is an atmospheric window where the longwave band light is not absorbed or scattered. This is the way that the Earth and the Atmosphere can "leak" energy back into space. If you like, go back to take a look at a similar figure from Granqvist. From which side of the sky window is water vapor absorbing the longwave light, and from which side is CO2?
Using this example, when incident light (irradiance) is not absorbed or transmitted through the surface and bulk of a material, it is reflected by the surface (the fraction of reflection is called the albedo).
I would like you to review Figures 5.8 and 5.9 in SECS Chapter 5 relating to the spectral reflectances of the following substances across a broad range of wavelengths:
Material properties that interact with light from the Sun (terrestrial shortwave band is around 250-2500 nm) extend over a far greater spectrum than what our eyes can perceive (visible sub-band 380-780 nm). Consider the shortwave range of light, 250-2500 nm. Anything with an albedo (reflectance) >0.2 could be a good contributor to a diffuse ground reflectance that increases light incident on a tilted collector. Even if the reflectance only cover a portion of the full spectrum from the Sun. Natural surfaces and built surfaces can be used to promote or inhibit solar reflectance, coupled to the performance of a Solar Energy Conversion System. Keep that in mind in the future!
Question: Which of these objects would be effective reflecting surfaces for shortwave irradiance? Which surface surprises you in its behavior?
There is a common phrase used in the solar world, which holds two distinct meanings to two user bases. First, I want you to understand the difference between an air mass coefficient used to qualify the testing conditions of a PV cell, and then I want you to understand the air mass (relative to source regions) that we will use to help in our assessment of the dynamic solar resource for clients in their multiple climate regimes.
For the good of the engineering community, air masses were "fixed" in 1976 as the U.S. Standard Atmosphere [16] for the ideal clear sky atmosphere in the mid-latitudes of the USA. Yep, and nothing has changed since 1976, right? (cough)
Actually, the American Society for Testing and Materials (ASTM) developed the "Terrestrial Reference Spectra for Photovoltaic Performance Evaluation" so that we could compare and evaluate the performance of our solar technologies, like PV cells and modules. The full documentation of the reference standard is held by the National Renewable Energy Laboratory (NREL AM Standard [17]).
The specified atmospheric conditions are:
Actually, we are speaking of an air mass coefficient: a relative measure of the optical path length passing through the Earth's atmosphere [19], as described for a fixed moment in time and space. We express the air mass coefficient as a ratio of the direct path of the global shortwave irradiance incident upon a specially tilted surface, relative to the path length for a horizontal surface (optical path oriented vertically upwards as the normal).
The first thing we need to estimate is what the intensity of the global horizontal irradiance would be just "outside" of our thin covering atmosphere. So, how do we determine the solar energy incident upon the extraterrestrial surface for the Earth-Atmosphere system?
First, we begin with the annual average solar constant, in units of irradiance (recently updated as of 2012! No longer 1366 or 1367 W/m2 if you knew that value). The average annual Solar intensity will actually vary according to an oscillation linked to 11-year cycles of sunspots. Given a large number of sunspots, the solar constant will be higher (~ 1362 W/m2) while the value will drop to ~ 1360 W/m2 when there are not many sunspots (changes about 0.01%). You can see NASA's Solar Radiation Climate Research Area [20] (accessed Feb 13, 2013).
Here, G is for Global irradiance, and "sc" stands for solar constant.
Second, we estimate the irradiance of the Sun for the intensity collected on a plane perpendicular to the beam shooting out from the Sun that is also perpendicular to the surface of the Sun (the normal radiant exitance). We would term that "normal global irradiance at AM0" (G0,n). This would be the equivalent of an estimate at zero cosine projection error, or the angle of incidence is zero. The cosine function in the following Equation delivers the change from zero to one, reflecting the cyclic orbit of the Earth relative to the Sun. The parameter n in this equation is the day number in the year.
Here, G is for Global, subscript "0" is for Air Mass Zero, and subscript "n" stands for radiant exitance normal to the surface of the Sun (don't confuse it with the day number).
During the Australian summer, the actual value exceeds 1416 W/m2, and then drops down to around 1326 W/m2 during the summer in the Northern Hemisphere.
Why would a Canadian summer not be as bright as an Australian summer (in outer space, of course)?
Third and finally, we need to estimate the AM0 irradiance for any given horizontal surface on Earth's surface (G0). Think about it, a horizontal surface on Earth (or just outside the atmosphere) will be tangent to the curvature of that locale on Earth. We should notice in the following equation that the Earth-Sun relations of declination ( ), latitude (), and hour angle () are all included in this irradiance equation for AMO.
By comparison with our work from Lesson 2, I can point out that the equation is actually multiplying the Equation for G0,n with the Equation for the solar altitude angle: Another way to write this would be Wow! Lesson 2 really comes in handy!
As a reminder, the relation to find is:
Consider that the sine of the altitude angle ( ) is the same as the cosine of the zenith angle ( ).
Other equations of use: The notation for irradiance is G, but the notation for irradiation (energy density, in J/m2 or Wh/m2) is broken down into hourly values and daily values. In the standard literature, hourly values use the coefficient I, while daily values use the coefficient H (no, I don't know why--it's just a quirk).
where the hour angle for sunset has been defined in Lesson 2 as:
The "standard" for testing. The air mass coefficient is defined in proportion to the cosine of the zenith angle (): the angle between the beam from the Sun and the normal vector pointing directly up to the zenith of the sky (normal to the horizontal surface).
One needs
By equivalence, the cosine of the zenith angle is the sine of the altitude angle , as we just demonstrated above.
Anecdotally, we noticed that the "ideal, engineered" clear sky conditions for AM1.5 or better (<AM1.5) are related to a window of time in a given locale. As a rule of thumb in the industry, we often mark off approximately 6 hours (centered on solar noon) for "prime" solar resource times during the day.
If one were to invert the problem and instead calculate the hour angle (and thus the time) from the air mass coefficient and zenith angles, we would find a mirror image of two times when that AM1.5 condition occurs on an ideal clear day in the summer (near the solstice). In the mid-latitudes of the continental USA or Europe, those instances of AM1.5 would occur sometime before 9am and after 3pm solar time. Any time within that window, the air mass coefficients would be less than 1.5.
Voila, the six-hour rule of thumb emerges from the Air Mass calculations.
Figure 3.9 illustrates the concept of airmass with respect to a zenith angle of (AM1.5), or an altitude angle of The hypothetical receiving surface would be tilted at ß = 37º for global irradiance Gt and an angle of incidence of
Gueymard has since developed a modified air mass model that will also apply for zenith angles >80°, as the former model tends toward infinity at .
Source: Matthew J. Reno, Clifford W. Hansen, and Joshua S. Stein. Global horizontal irradiance clear sky models: Implementation and analysis. Technical Report SAND2012-2389, Sandia National Laboratories, Albuquerque, New Mexico 87185 and Livermore, California 94550, March 2012.
Why is meteorology so important to solar energy conversion systems? Think about meteorology and the scientists that explore the field. In effect, they can help us in the analogous manner that a geologist helps to define and quantify mineral resources in the ground. It is up to us to learn the language of meteorology to communicate effectively with our peers and increase our success in project development and planning.
As noted in the text book, we need to use the Air Mass phrase suitable for the systems science perspective, reflecting changing parcels of air shifting across the surface of Earth. Those parcels are generating clouds and dust storms, tornadoes, hurricanes, and sunny blue skies. This is the meteorological concept of the Air Mass that we will use to help in our assessment of the dynamic solar resource for a client in their multiple climate regimes.
The Air Masses used in engineering standardization for SECS technologies are not nearly as useful as the real, dynamic air masses that we will use for solar resource assessment.
The sky can be divided up into large volumes with common properties. Air masses are large pancakes of gases and particles (we call them parcels) with common properties of temperature, chemistry, and pressure. Aside from the strong role of clouds, those common physical properties each affect the way in which solar energy intensity will be decreased along the particular path length through the sky. Air mass will affect the way that light is absorbed and scattered, and hence will affect the characteristic behavior of light incident upon an aperture. The changes between air masses are weather fronts, and the study of the regional dynamic dance of air masses and accompanying fronts on a sub-annual basis is called meteorology. When we expand our studies of air mass behaviors to multiregional scales (larger spatial scales) and time scales in the range of decades to millennia, this is called climatology.
I want you to think of an air mass from meteorology as a turbulent pancake that interacts with both shortwave and longwave light, resulting in highly variable irradiance conditions at the ground level, and also for satellite remote sensing in orbit. Irradiance conditions are constantly shifting over a given locale, and we will be in need of effective forecasts of irradiation (over different time intervals) for many SECS technologies.
It is always important to solar energy conversion that we figure time and scale into our understanding. We observe that each location under study is influenced by multiple air masses during different times of the year. In the midlatitudes of North America, we call this periodicity seasons (deep thought, yes). Hence, our solar energy conversion systems are effectively in different locales for each period of air mass dominance.
One locale for your client will be divided up into seasons. For us, each season (a block of time) is like an independent region, different from the next season--and we call these regions climate regimes. In the mid-latitudes we observe four climate regimes, while regions affected by monsoonal swings may have two or three climate regimes.
Consider: there is not one State College, but four! One State College for each season, or specifically, one statistically different solar resource fingerprint for each synoptic climate regime.
Regime changes are often noticed by us in terms of the air temperature and humidity, but the sky regimes will change as well, with respect to wind speed, the size of weather cells, and emergent cloud behaviors.
From the perspective of design in solar energy conversion systems, the skies allowing sunlight through to Philadelphia from December 1 through February 28 exist in a different climate regime from the Philadelphia skies of June 1 through August 31.
Recall that the change of irradiation over seasons is due to the tilt of Earth (which leads to our measure of declination). When the Earth is unevenly energized by the Sun, air masses are mobile. Any locale will observe changes in residential air mass behavior over time. Air masses are created in source regions, where the mass acquires its dominant characteristics of temperature and humidity. Then air masses move from the region of origin into new regions. Source regions have been mapped out in Figure 3.11, below, and they will tend to have lighter winds, allowing the air mass to accumulate the temperature and humidity conditions of the accompanying portion of the Earth's surface (the big solar energy conversion device). Hence, areas affected by jet streams will not be regions to create new air masses.
The Bergeron classification system is accepted and used by atmospheric science and indicates the origin locale. This is important information, affecting the quality of the solar resource. Properties of thermal behavior and humidity are conveyed in the Bergeron classification, and the classification has a relatively simple two or three letter coding.
First, air masses are labeled according to their origin above a land mass, continental (c) or above an ocean, maritime (m). Maritime air masses will contain more humidity derived from the underlying body of water. Next, air masses can be grouped into main thermal categories of the site origin: (A, P, T, E). This grouping is largely separated by jet streams and changes in latitude, which of course are each derived by the tilt of Earth relative to the projection of irradiance from the Sun. By coupling the two letters together, we can then map out the source regions for meteorological behaviors across the planet and through the seasonal shifts.
The relation of air mass to monsoon cycles is less used in continental meteorology of North America, but is highly important to solar development in Asia.
Again, air masses are labeled according to the origin of the parcel above a land mass,
Air mass source regions are also grouped according to their thermal characteristics, the thermal generator that is charging up the parcel:
Clouds have the ability to scatter and even focus light, which means that sometimes a cloud will remove energy from the solar resource on the ground, and other times clouds will collect and focus light that's even more intense than light on the clearest day in your locale. Yet, we know that clouds are ephemeral and dynamic systems in and of themselves.
The albedo is the reflective fraction of light from a given surface, a fractional value from 0-1.
Let's consider a jar of clean water (upper left in the photograph), into which one may add a small amount of milk (upper right). Milk is a "colloidal suspension" of solid particles supported/floating in liquid. As such, the solid particles of fat reflect light, scattering the light in many directions like a cloud would do.
In the lower left and lower right corners, the milk concentration has been steadily increased. We can see through the clear glass of water, suggesting that the transmission of visible light from behind the glass to an observer standing behind the camera is high. For the low concentration mixture of milk and water in the upper right, the small amount of added milk has slightly decreased the transmission of light. We can still see the louvers of the window blinds behind the jar, however.
With more milk added to the mixture in the glass on the lower left, the transmission of light was further reduced, and one's ability to perceive the louvers on the other side of the glass was diminished. Finally, adding more milk to the glass in the lower right cut the transmission of visible light to very low levels, and we can't see through the glass.
What is not being transmitted or absorbed is being reflected. Consider this experiment in the context of albedo. The albedo of clean water in the visible sub-band is much lower than the increasing albedos of all three mixtures of water and milk. Of the three mixtures, the glass with the greatest concentration of milk has the highest albedo.
Recap: we have seen how reflection and albedo is a bit more complicated when light is interacting with millions of surfaces in a suspension, like a cloud. We can also see why a more effective term is backscattered radiation, which is used to describe the collective phenomena. And yet, we can use the backscattered data from clouds to infer useful information about the solar resource at the ground. How do you think we could use that data?
Time to explore the concept of climate regimes and the influence of meteorological fingerprints in solar resource assessment.
One locale for your client will be divided up into seasons. For us, each season (a block of time) is like an independent region, different from the next season--regions that we will call climate regimes. In the mid-latitudes, we observe four climate regimes, while regions affected by monsoonal swings may have two or three climate regimes.
Recall: there is not one St. Louis, there are four! One for each season, or one statistically different solar resource fingerprint for each synoptic climate regime of St. Louis, Missouri.
Climate regime changes are often noticed by us in terms of the air temperature and humidity, but the sky regimes will change as well, with respect to wind speed, the size of weather cells, and emergent cloud behaviors.
I want you to think about the way that seasonal (or synoptic) variations in a locale will generate multiple fingerprints. While each fingerprint is different from another in a given locale (translated: winter solar is not like summer solar), additionally, each set of fingerprints is different from another regional set (translated Mumbai monsoons are not like humid summers in Missouri). Every one of our locales for solar design is going to have "cloudy conditions" at some time, and everyone will have "sunny days" at other times. How can we better understand the variety of intermittencies and trends for our locale and synoptic climate regime (our fingerprint)?
This is just a play on words to emphasize something that you probably can guess: these three locations are not like each other!
Please identify the contributions or physical parameters that relate to the solar resource (other than the sun). First look at the questions and write down your answers in your notes. Then scroll over the answers to see how your thoughts match up.
1. What is the role of the sky dome in the solar resource?
Click for answer.
2. What is the role of clouds in the solar resource?
Click for answer.
3. What is the role of aerosols in the solar resource?
Click for answer.
We just introduced a totally foreign (yet intuitive) concept of climate regime to our lexicon in solar design. This is a new concept, but an essential one for advanced project development. You can start the discussion by posting what you perceive the meaning of climate regime to be from your personal reading and thought process. You can also comment on any uncertainty you initially found regarding the meaning, the intent, or application of climate regime to a project design, or to long-term project management. Creative interpretations of potential implications are welcome!
Here are some guiding questions:
When you create a post in the Yellowdig discussion space, you are required to choose a topic tag. For Lesson 3 discussions, feel free to use any of the following:
You can tag your post with one or several topics at the same time. All posts and contributions you create are added up to one score at the end of the week.
Yellowdig tip: check in the Yellowdig site at least once per day. Commit to making at least one contribution daily – read new posts, ask questions, give your peers reactions and accolades. If reading generates some thoughts, share them, don’t postpone until later. This is a team learning space, so you are also helping others learn by being more active.
Yellowdig points you earn over the weekly point earning period (from Saturday to next Friday) will count towards 1000 pts. weekly target. But you can go above it (to 1350 pts. max). Yellowdig discussions will account for 15% of the total grade in the course. Check back the Orientation Yellowdig page in Canvas for more details on the points earning rules.
There is no hard deadline for participating in these discussions, but I encourage you to create your posts in the middle of the study week (Sunday) to allow others to engage and respond while we are learning specific topics in the lesson. Also, remember that each weekly point earning cycle ends Friday night, so all points accumulated by then will represent your weekly grade, and then you start all over again (it’s like playing a set in tennis..)
In the section discussing components of light, we reviewed the effects of the atmosphere on irradiance. Now, we discuss how to model the transparency of a hypothetical sky that is "clear" of the effects of clouds. Clouds are a major contributor to the reduction of terrestrial irradiance; however, particles in the atmosphere also play a significant role. There are physical properties of the sky that we cannot see with our eyes, but which strongly affect the solar resource on a clear day. As we will see in the reading, aerosols and water vapor present in the atmosphere play an important role in scattering light, and they may be present on "clear sky" days when visible clouds are absent.
Daily or monthly irradiance data is required for proper design of any solar energy collection system. However, this data is not always available. This requires the use of well-designed models to estimate irradiance. Hence, the need for clear sky models.
Clear sky models are used to estimate what is called a clearness index. For a location, a clear-sky model must be properly calibrated to provide an accurate measure for the clearness index. We will be looking at two modeling approaches, the Bird Clear Sky Model (which we can download and use as a spreadsheet) and the REST2 model by Gueymard (which can also be downloaded and run as an executable file--we will not do so here).
The Bird Clear Sky Model was developed by Richard Bird and a number of other scientists at what is now the Department of Energy National Renewable Energy Laboratory. The model requires the following input data
Many openly available codes incorporate this model. The output is a "clear sky" estimate for the total or global horizontal irradiance (GHI), direct normal irradiance (DNI) and diffuse horizontal irradiance (DHI, or DIF) across wavelengths from 305nm to 4000nm. The model calculates these conditions for a single point in solar time, given the latitude (), longitude () and the Time Zone.
Download RReDC: Bird Clear Sky Model [29]
Note: You will be asked to use this tool in this week homework assignment.The REST2 model has been found to be most accurate, as we shall observe in our reading. We will only need to explore one method (Bird) for this class, but it is important that a resource professional is also aware of the modern application of clear sky modeling. REST2 accepts atmospheric inputs of:
The REST2 model will then output estimations of diffuse horizontal irradiance (DHI), direct normal irradiance (DNI), and global plane of array (POA) irradiance.
Clouds are the big challenge in solar resource assessment. Think about how we just established a "clear sky" model, but most of our skies are filled with clouds that are dynamic and diverse in character. What is it about clouds that makes resource assessment challenging? Take a moment now to familiarize yourself with cloud phenomena and types.
Clouds are emergent phenomena within the atmosphere, and strongly perturb the behavior of the solar resource in a given locale. When we do not have a clear sky day in our locale (typical in many locations on Earth), the effects of clouds will strongly draw down the beam component of our solar resource and increase the net fraction of diffuse sky irradiance.
Clouds can develop in environments that are termed active and passive, as well as from updrafts near the Earth’s surface. As we have seen in the reading, clouds have the ability to either scatter and reduce incident light, or to act as a lens at the perimeter, focusing light well above clear sky or even AM0 irradiance conditions.
From the perspective of an observer standing on the Earth's surface, clouds can be classified by their physical appearance. Accordingly, there are essentially three basic cloud types:
Let's review some basic types of clouds. Please do comparative reading between the figure below (with images) and your list of cumuliform, stratocumulus, and stratiform clouds in Ch. 5 of SECS.
Clouds are listed here with images that pop up when you click on the red dots. We see variants of cumuliform clouds, stratocumulus clouds, and stratiform clouds, as noted in our reading.
There are four general classifications of clouds: high, middle, and low clouds as well as clouds of vertical development. The table below summarizes each of these classifications while giving you a sense for the typical altitudes at which their cloud ceilings (the height of their bases) are observed, and the basic chemistry of the clouds (ice and water interact with light in different ways).
Recapping: You should begin to think about how clouds can both block the beam component of light from the Sun (making the skydome more diffuse in nature), and can refract and scatter light like a lens, increasing the irradiance on a locale far higher than predicted from a clear sky model.
Researchers are now trying to predict cloud behavior with respect to GHI and DNI variation in a given locale. We are still a few years off, but soon we may see solar forecasts that include cloud interference for solar resource assessment.
Acknowledgment: Content of this page comes from Meteo 101: Understanding Weather Forecasting; author: Lee Grenci and David Babb.
SECS, Chapter 5: Meteorology, "Robot Monkey Does Space-Time" section
When we look at clouds, think of how they float by us without really changing in form all that much. Kind of like train cars passing by on the railroad, only much slower. We can use this to connect space and time (or frequency) in a useful way for solar resource assessment.Clouds (and weather fronts) occur on multiple scales of space and time. These scales contribute to the concept of solar fingerprints or climate regimes. In our reading, we learn about Sir Geoffrey I. Taylor and his work on turbulence theory. Clouds lead to intermittent solar conditions (POA and DNI irradiance) that, in turn, affect our solar energy conversion technologies, such as PV and building systems.
The intermittent variation of clouds as they affect the solar resource is crucial when solar is deployed as a large-scale farm, or as a large community of decentralized PV arrays on buildings. Variable peaks and valleys of electrical power can be detrimental (costly) to utilities if there are no other ways to store or shave variation on the grid. Now, storage solutions are emerging slowly in research, but right now we need to address the intermittency problem by better understanding the drivers and the basic science underneath those drivers.
A series of changes in time at a fixed place is due to the passage of an unchanging spatial pattern over that locale. That is, when observing a cloud or thunderstorm passing overhead, the clouds of the thunderstorm floating by are effectively "unchanging spatial patterns" (big blocks of cloud). Put another way, the lateral change in the cloud conditions across regions of the meteorological event (e.g., cloud-sky-cloud-sky-cloud) can be directly connected locally with a variation in irradiance measurement over time (e.g., a periodic measure of dark-bright-dark-bright-dark).
Taylor’s hypothesis states that events that change in time for a fixed place are due to the flow or passage of unchanging spatial patterns over a locale. This is like observing a thunderstorm passing by directly overhead and making the connection to the rate of the clouds advecting with the thunderstorm. Taylor's hypothesis from the 1930s allows us to flip from the Eulerian frame of reference (time sequences) to the Lagrangian frame of reference (spatial changes).
The Eulerian frame of reference: When an observer is stuck in one place, only watching the changing phenomena as it passes by.
The Langrangian frame of reference: When the observer moves with the meteorological phenomena instead of remaining fixed. Imagine a flying carpet floating along with a cumulus cloud moving from one county to the next.
Hence, all scales of time (or frequency) are also spatial scales! Taylor's Hypothesis holds so long as the advective wind speed is much greater than the timescale of the evolving meteorological event being investigated, as is often the case. When you have a lazy cloud day, where the clouds are changing form faster than they move, the time-space connection doesn't hold anymore.
By using spatio-temporal scales of cloud features established by Ted Fujita, we can estimate an average translation (advection) speed of 17 m/s for meteorological phenomena. We can then convert the spatial scales of variability into a relevant timescale. Spatial scales associated with power transmission congestion are distances of 25-1000 km. These distances are relevant for meteorological phenomena within time scales on the order of 25 seconds to 16 hours.
Cumulus 2-5 km, which means 10-100 minutes.
Cumulonimbus (including anvil) 10-200 km, which means 1-5 hours.
Cumulonimbus cluster (including merged anvils) 50-1000 km, which means 3-36 hours.
Synoptic (including cyclone waves, short and long Rossby waves) 1000-40000 km, which means 2-15+ days.
If we plot all of these phenomena and draw a line down the middle, we have a really rough estimate of the average advection rate of meteorological phenomena on Earth. Here, we have represented that rate in three different unit scales:
As such, we can connect Synoptic, Mesoscale, and Microscale meteorological phenomena between spatial and temporal scales!
Synoptic Weather: This is weather on the scale of 1000+ km in distance, and (given advection of ~17 m/s) seasonal timescales of 2-15+ days.
Mesoscale Weather: This is weather phenomena, including solar condition, on the scale of 100-1000 km distances, and time scales less than one day (hours).
Microscale Weather: This is variable weather phenomena on the scale of 10-100 km distances, and time scales of minutes to seconds.
Taylor’s Hypothesis “works” when the air in the sky moves significantly faster (advection: wind speeds) than the evolution time of a cloud or a thunderstorm of clouds. So, this approximation breaks down for those days when the cumulus clouds just "hang" in the sky all day.
The following readings will be helpful for completing this lesson activity.
Especially:
The goal of this activity is to use the BIRD Clear Sky Model to predict the solar irradiance at a specific time at a specific locale and further compare those predictions to actual irradiance measurements.
So there are two main tasks in this activity:
These two tasks will be performed for a specific date - July 31, 2007 (in this assignment) for the specific locale - SURFRAD Meteorological Station at Rock Springs, located just outside State College, PA.
In the end, you will need to assess how the modeled (predicted) and measured data fit and provide some discussion on it.
The Bird model is an old but advanced clear sky model used by solar professionals to estimate clear sky conditions at a locale. The clear sky model for a horizontal surface is the basis for almost all modern comparisons of "ideal skies" to actual skies (with clouds and dynamic light interference effects). You will see what the global, beam horizontal, and diffuse components of solar light should look like when you set the meteorological parameters affecting the sky clearness. The model could also provide DNI (Direct Normal Irradiance), which would show irradiation values higher than the global horizontal values for the mornings and evenings.
Note: access the AOD data through this graphical page [36]. Take the daily average as input for the Bird model. In this case, 415nm AOD can be used as a proxy for 380nm AOD input and 500nm AOD data can be used as is.
The model plot will look similar to the one below. This is what you need to produce.
The SURFRAD site at Penn State collects real-time data for solar irradiance (using local pyranometers) and a number of other atmospheric parameters at 3-minute time intervals. These data are accessible through the Earth System Research Laboratory website [37]. This site provides you with access to data collected at the Penn State location. For each year, you can open the parent directory with .dat files, which are named by the day number. So if you are looking for July 31st, for example, that will be day #212, so check the file psu07212.dat.
Your goal will be to make sense of these data, extract your irradiance values and plot them versus time of the day.
Here are the data labels in order that correspond to the numbers in the .dat file (omit 0-value columns):
Year, jday, month, day, hour, min, dt (decimal time), zen (degrees), dw_solar (W/m^2), uw_solar (W/m^2), direct_n (W/m^2), diffuse (W/m^2), dw_ir (W/m^2), dw_casetemp (K), dw_dometemp (K), uw_ir (W/m^2), uw_casetemp (K), uw_dometemp (K), uvb (mW/m^2), par (W/m^2), netsolar (W/m^2), netir (W/m^2), totalnet (W/m^2), temp (degC), rh (%), windspd (m/s), "winddir (degrees, clockwise from north)", pressure (mb)
From this list, you will specifically need:
dt (decimal time) - your time coordinate for data plotting (Column G in MS Excel)
dw_solar - Downwelling global solar (W/m^2) = corresponding to Global Hz (GHI) in BIRD model (Column I in MS Excel)
netsolar - Net solar (W/m^2) = corresponding to Direct Hz (DHI) in BIRD model (Column AG in MS Excel)
diffuse - Real downwelling diffuse solar (W/m^2) = corresponding to Dif Hz (DIF) in BIRD model (Column O in MS Excel)
Also, 'pressure (mb)' - the last value in each array of data - daily average should be used for your Bird model input (Column AU in MS Excel).
To treat the data, you can either convert the .dat file to a spreadsheet (see video in Canvas Module 3 how to!) or, if you have programming skills, you can create a script to plot the irradiance data versus time.
Important Note: you may notice that the "0" decimal time in the .dat file does not correspond to the actual local midnight, but rather corresponds to Greenwich 0:00, so you may need to synchronize decimal time with the Bird model to make the calculated and measured curves match.
Prepare a written report including the following:
Save your report as .docx or PDF and submit to the Lesson 3 Activity - Clear Sky Model dropbox in Canvas.
Please see the grading rubric in Canvas
See the Calendar tab in Canvas for specific due dates.
In this lesson, we have learned about the diverse ways in which the atmosphere can interact with shortwave light to affect the solar resource received at the ground by our Solar Energy Conversion Systems. We have observed that the sky has a distinct character derived from air mass source regions, and that the weather has seasonal "fingerprints" that distinguish a single site as multiple unique climate regimes. We further found that a "clear sky" is a fairly complex system of its own, and modeling the clear sky is not trivial. When we finally add clouds to the skies, we see where the real source of solar intermittency stems from. By connecting space and time from Taylor's Hypothesis, we find that we hold additional information on the scales of intermittency to expect in a given locale and for each seasonal climate regime in that same locale.
You have reached the end of Lesson 3! Double-check the to-do list on the Lesson 3 Learning Outcomes page to make sure you have completed all of the activities listed there before you begin Lesson 4.
Links
[1] https://solardata.uoregon.edu/download/Papers/SolarCellBasedPyranometers--EvaluationoftheDiffuseResponse.pdf
[2] http://www.ssd.noaa.gov/goes/east/
[3] http://www.ssd.noaa.gov/goes/west/
[4] https://pmm.nasa.gov/resources/glossary#letter_r
[5] http://www.sciencedirect.com/science/article/pii/S0038092X11004221
[6] http://www.sciencedirect.com/science/article/pii/S0038092X13000947
[7] https://www.e-education.psu.edu/eme810/node/623
[8] https://creativecommons.org/licenses/by-nc-sa/4.0/
[9] http://hyperphysics.phy-astr.gsu.edu/hbase/vision/isql.html
[10] http://en.wikipedia.org/wiki/File:Inverse_square_law.svg
[11] https://creativecommons.org/licenses/by-sa/3.0/deed.en
[12] http://www.wolframalpha.com/
[13] http://www.wolframalpha.com
[14] https://commons.wikimedia.org/wiki/File:Atmospheric_Transmission.png
[15] https://www.e-education.psu.edu/eme810/sites/www.e-education.psu.edu.eme810/files/images/Lesson_01/backscatter.png
[16] http://www.pdas.com/atmos.html
[17] http://rredc.nrel.gov/solar/spectra/am1.5/
[18] http://speclib.jpl.nasa.gov
[19] http://en.wikipedia.org/wiki/Earth%27s_atmosphere
[20] https://earth.gsfc.nasa.gov/climate
[21] http://www.nrel.gov/grid/solar-resource/renewable-resource-data.html
[22] https://www.nrel.gov/
[23] http://commons.wikimedia.org/wiki/File:Worldmap_wdb_combined.svg
[24] http://en.wikipedia.org/wiki/User:gringer?rdfrom=commons:User:Gringer
[25] http://en.wikipedia.org/wiki/St._Louis#Climate
[26] http://en.wikipedia.org/wiki/Honolulu#Climate
[27] http://en.wikipedia.org/wiki/Climate_of_Mumbai
[28] https://www.nrel.gov/docs/legosti/old/761.pdf
[29] http://rredc.nrel.gov/solar/models/clearsky/
[30] http://en.wikipedia.org/wiki/Angstrom_exponent
[31] http://earthobservatory.nasa.gov/GlobalMaps/view.php?d1=MODAL2_M_AER_OD
[32] https://www.e-education.psu.edu/eme810/sites/www.e-education.psu.edu.eme810/files/images/Lesson_03/Ci_3.jpg
[33] https://www.e-education.psu.edu/eme810/sites/www.e-education.psu.edu.eme810/files/images/Lesson_03/Acumulus.jpg
[34] https://www.e-education.psu.edu/eme810/sites/www.e-education.psu.edu.eme810/files/images/Lesson_03/stratus3.jpg
[35] https://www.e-education.psu.edu/eme810/sites/www.e-education.psu.edu.eme810/files/images/Lesson_03/cu_congestus.jpg
[36] http://www.esrl.noaa.gov/gmd/grad/surfrad/aod/aod_check.php?site=psu&date=2007-07-31&wavelength=415
[37] https://www.esrl.noaa.gov/gmd/dv/data/index.php?category=Radiation&parameter_name=Surface%2BRadiation&site=PSU