أرشيف المدونة الإلكترونية

صور المظاهر بواسطة MichaelJay. يتم التشغيل بواسطة Blogger.
أقسام المدونة :

الخميس، 28 يناير 2016

6th Asian Mining Congress and Exhibition - 2016


AMC 2016
Asia, the world's biggest continent, is also the world's largest raw material producer, to the tune of 10 billion tonnes per year, accounting for 56% of world total mineral production. With its huge area and geological diversity, it is likely and virtually every kind of deposit is present somewhere. The outlook for development of the mineral industry is promising and it will play a major role in the global economy to achieve its full potential.

The Asian Mining Congress and International Mining Exhibition, a biennial event organized by the Mining, Geological & Metallurgical Institute of India (MGMI) provides a forum for the miners, machinery manufacturers, planners and policy makers to discuss the various issues affecting the mining industry in the Asian region in particular, and also in the rest of the world. The event provides an unrivalled opportunity for the manufacturers of mining machinery in the world to showcase their products and do business. 

The Congress will provide forum for promotion and support of techno-scientific cooperation towards national and international progress in the areas of mineral production, in addition to the development of new opportunities of sustainable business that will benefit both Asian and world societies.

Lead Topics
  • Status of Mineral Industry in Asian Countries: Resources and Exploitation.
  • New Mineral Development Projects in Asian Countries.
  • Oil and Gas: Petroleum, Natural Gas, Coal Bed Methane (CBM), Coal Mine Methane (CMM), Shale Gas, Underground Gasification of Coal (UGC), Coal Liquefaction, etc.
  • Advances in Technology: Exploration & Mining in Opencast & Underground Modern Mining Techniques: Opencast and underground Mining.
  • Mineral Processing and Coal Beneficiation.
  • Environment, Safety and Health issues.
  • Investment opportunities in Mining Industry.
  • Green Mining for sustainable future.
  • Import, Export Trading scenario.
  • Logistics & Infrastructure Development.

Highlights of International Mining Exhibition
  • Participation of leading Mining Equipment & Machinery Manufacturers from Asia, Africa, America and Europe.
  • Group Participation from Mineral Rich Countries and States of India.
  • Participation of Industry Giants and SMEs'.
  • Structured Visitor Promotion and Publicity to Invite High Purchase Buyers.
  • Buyer Seller Meet.
  • Participation expected from 30+ Countries, 300+ Stalls, 800+ Delegates, 20,000+                Trade Visitors.

Conference Date: 23rd – 27th February, 2016.
Conference Venue: Hotel Hyatt Regency, Kolkata, West Bengal, India.
Exhibition Venue: Eco Park (Gate No.1), New Town, Rajarhat, Kolkata, WB, India.
Contact: The Mining, Geological & Metallurgical Institute of India, GN - 38/4, Salt lake, Sector – V, Kolkata – 700 091, India.
Website: AsianMining Congress  ; MGMI             
Email: mgmikolkata@gmail.com; info@asianminingcongress.com Tel: 91 33 2357 3482/3987/6518


.


Kumba Iron Ore Ltd begins workforce reduction process

Sishen Iron ore Mine

Kumba Iron Ore, South Africa’s largest producer of the crucial steel ingredient, is slashing the workforce at its flagship Sishen mine by nearly half to cope with weak iron ore prices.

The Sishen mine, the largest source of iron ore in South Africa for decades, has undergone a major change because of the enormous amount of waste that had to be moved to expose ore.

The change has not been quick enough and majority shareholder Anglo American told investors last month a decision had been taken to shift focus away from volumes and instead focus on cutting costs, reducing capital expenditure and boosting cash generation.

The production forecast was lowered to 26-million tonnes a year at a unit cash cost delivered on board ships in Saldanha port of less than $30 a tonne this year, giving a break-even price of landing ore in China at below $40 a tonne.

As part of this new plan, the workforce would have to be cut, Kumba said yesterday. Of the 5,840 employees at Sishen, Kumba aims to reduce the number by cutting 2,633 direct employees and 1,300 contractor jobs.

Kumba issued the unions a section 189 notice on Thursday, starting a 60-day process to reduce the workforce.

“This has been an extremely difficult decision but, after exhausting all other avenues and doing all we could have done to reduce costs, we have no choice but to take more significant steps to preserve the viability of the mine,” CEO Norman Mbazima said.

Sishen is the largest single source of jobs in Kumba, which employs 7,434 people after it stopped mining the Thabazimbi mine last year, removing 1,160 jobs. In July last year, Kumba told the unions it wanted to cut 175 jobs at its Northern Cape mines — Sishen and Kolomela.

“It cannot be correct that as and when the mining industry is under distress the first casualties are ordinary employees,” said Lucas Phiri, National Union of Mineworkers chief negotiator at Kumba.

The dramatic fall in the iron-ore price, which Bloomberg pegged at less than a quarter of its peak in 2011 — hovering at the $41 a tonne mark — has put pressure on the iron ore producer’s balance sheet.

Kumba on Thursday reported a 12% drop in its fourth-quarter output to 10.04-million tonnes, dragged down by poor performance at Sishen. For the year, its output was down 7% at 44.88-million tonnes.

The Sishen mine was moving to a lower-cost pit layout and there was not enough exposed ore as the transition was being made.

Sishen’s output in the fourth quarter fell 17% to 7.7-million tonnes.

Kumba’s exports sales fell 10% to 10.5-million tonnes in the quarter and the company had a 4.7-million tonne stockpile, down from 6.5-million tonnes at the end of 2014.

Source: bdlive

What Are Earth Layers Made Of?

What Are Earth Layers Made Of? 

A modern view of Earth‘s interior layers.
As a result of studies during the past century, geologists have a pretty clear sense of what the layers inside the Earth are made of. Let’s now look at the properties of individual layers in more detail (figure above a, b).

The Crust 

When you stand on the surface of the Earth, you are standing on top of its outermost layer, the crust. The crust is our home and the source of all our resources. How thick is this all important layer? Or, in other words, what is the depth to the crust-mantle boundary? An answer came from the studies of Andrija Mohorovicˇic´, a researcher working in Zagreb, Croatia. In 1909, he discovered that the velocity of earthquake waves suddenly increased at a depth of tens of kilometres beneath the Earth’s surface, and he suggested that this increase was caused by an abrupt change in the properties of rock. Later studies showed that this change can be found most everywhere around our planet, though it occurs at different depths in different locations. Specifically, it’s deeper beneath continents than beneath oceans. Geologists now consider the change to define the base of the crust, and they refer to it as the Moho in Mohorovicˇic´’s honour. The relatively shallow depth of the Moho (7 to 70 km, depending on location) as compared to the radius of the Earth (6,371 km) emphasizes that the crust is very thin indeed. In fact, the crust is only about 0.1% to 1.0% of the Earth’s radius, so if the Earth were the size of a balloon, the crust would be about the thickness of the balloon’s skin.
The crust is not simply cooled mantle, like the skin on chocolate pudding, but rather consists of a variety of rocks that differ in composition (chemical make-up) from mantle rock. Geologists distinguish between two fundamentally different types of crust oceanic crust, which underlies the sea floor, and continental crust, which underlies continents. 
Oceanic crust is only 7 to 10 km thick. At highway speeds (100 km per hour), you could drive a distance equal to the thickness of the oceanic crust in about five minutes. At the top, we find a blanket of sediment, generally less than 1 km thick, composed of clay and tiny shells that settled like snow out of the sea. Beneath this blanket, the oceanic crust consists of a layer of basalt and, below that, a layer of gabbro. 

 A table and a graph illustrating the abundance of elements in the Earth’s crust.
Most continental crust is about 35 to 40 km thick about four to five times the thickness of oceanic crust but its thickness varies significantly. In some places, continental crust has been stretched and thinned so it’s only 25 km from the surface to the Moho, and in some places, the crust has been crumpled and thickened to become up to 70 km thick. In contrast to oceanic crust, continental crust contains a  great variety of rock types, ranging from mafic to felsic in composition. On average, upper continental crust is less mafic than oceanic crust it has a felsic (granite-like) to intermediate composition so continental crust overall is less dense than oceanic crust. Notably, oxygen is the most abundant  element in the crust (figure above).

The Mantle 

The mantle of the Earth forms a 2,885-km-thick layer surrounding the core. In terms of volume, it is the largest part of the Earth. In contrast to the crust, the mantle consists entirely of an ultramafic (dark and dense) rock called peridotite. This means that peridotite, though rare at the Earth’s surface, is actually the most abundant rock in our planet! Researchers have found that earthquake-wave velocity changes at a depth of 400 km and again at a depth of 660 km in the mantle. Based on this observation, they divide the mantle into two sublayers: the upper mantle, down to a depth of 660 km, and the lower mantle, from 660 km down to 2,900 km. The transition zone is the interval between 400 km and 660 km deep. 
Almost all of the mantle is solid rock. But even though it’s solid, mantle rock below a depth of 100 to 150 km is so hot that it’s soft enough to flow. This flow, however, takes place extremely slowly at a rate of less than 15  cm a year. Soft here does not mean liquid; it simply means that over long periods of time mantle rock can change shape without breaking. We stated earlier that almost all of the mantle is solid. We used the word “almost” because up to a few percent of the mantle has melted. This melt occurs in films or bubbles between grains in the mantle at a depth of 100 to 200 km beneath the ocean floor. Although overall, the temperature of the mantle increases with depth, temperature also varies significantly with location even at the same depth. The warmer regions are less dense, while the cooler regions are denser. The distribution of warmer and cooler mantle indicates that the mantle convects like water in a simmering pot; warmer mantle is relatively buoyant and gradually flows upward, while cooler, denser mantle sinks.

The Core 

Early calculations suggested that the core had the same density as gold, so for many years people held the fanciful hope that vast riches lay at the heart of our planet. Alas, geologists eventually concluded that the core consists of a far less glamorous material, iron alloy (iron mixed with tiny amounts of other elements). Studies of seismic waves led geo scientists to divide the core into two parts, the outer core (between 2,900 and 5,155 km deep) and the inner core (from a depth of 5,155 km down to the Earth’s centre at 6,371 km). The outer core consists of liquid iron alloy. It can exist as a liquid because the temperature in the outer core is so high that even the great pressures squeezing the region cannot keep atoms locked into a solid framework. The iron alloy of the outer core can flow, and this flow generates Earth’s magnetic field. 
The inner core, with a radius of about 1,220 km, is a solid iron alloy that may reach a temperature of over 4,700°C. Even though it is hotter than the outer core, the inner core is a solid because it is deeper and is subjected to even greater pressure. The pressure keeps atoms locked together tightly in very dense materials.

The Lithosphere and the Asthenosphere 

So far, we have identified three major layers (crust, mantle, and core) inside the Earth that differ compositionally from each other. Earthquake waves travel at different velocities through these layers. An alternative way of thinking about Earth layers comes from studying the degree to which the material making up a layer can flow. In this context, we distinguish between rigid materials, which can bend or break but cannot flow, and plastic materials, which are relatively soft and can flow without breaking.

A block diagram of the lithosphere, emphasizing the difference between continental and oceanic lithosphere.
Geologists have determined that the outer 100 to 150  km of the Earth is relatively rigid. In other words, the Earth has an outer shell composed of rock that cannot flow easily. This outer layer is called the lithosphere, and it consists of the crust plus the uppermost, cooler part of the mantle. We refer to the portion of the mantle within the lithosphere as the lithospheric mantle. Note that the terms lithosphere and crust are not synonymous the crust is just the upper part of the lithosphere. The lithosphere lies on top of the asthenosphere, which is the portion of the mantle in which rock can flow. The boundary between the lithosphere and asthenosphere occurs where the temperature reaches about 1280°C, for at temperatures higher than this value mantle rock becomes soft enough to flow. 
Geologists distinguish between two types of lithosphere (figure above). Oceanic lithosphere, topped by oceanic crust, generally has a thickness of about 100 km. In contrast, continental lithosphere, topped by continental crust, generally has a thickness of about 150 km. Notice that the asthenosphere is entirely in the mantle and generally lies below a depth of 100 to 150 km. We can’t assign a specific depth to the base of the asthenosphere because all of the mantle below 150 km can flow, but for convenience, some geologists consider the base of the asthenosphere to be the top of the transition zone.
Credits: Stephen Marshak (Essentials of geology)

الأربعاء، 27 يناير 2016

Canadian Mining Journal January 2016


Canadian Mining Journal is the leading exploration and mining journal in Canada. Covers mineral exploration trends, Corporate Trends,  metal prices and new geological models, underground mine developments and operating performances, unique challenges of open pit operations and more. It is Canada's First Mining Publication.

الثلاثاء، 26 يناير 2016

Introducing the Earth’s Interior

Introducing the Earth’s Interior 

What Is the Earth Made Of? 

The proportions of major elements making up the mass of the whole Earth.
At this point, we leave our fantasy space voyage and turn our attention inward to the materials that make up the solid Earth, because we need to be aware of these before we can discuss the architecture of the Earth’s interior. Let’s begin by reiterating that the Earth consists mostly of elements produced by fusion reactions in stars and by supernova explosions. Only four elements (iron, oxygen, silicon, and magnesium) make up 91.2% of the Earth’s mass; the remaining 8.8% consists of the other 88 elements (figure above). The elements of the Earth comprise a great variety of materials. 
  • Organic chemicals. Carbon-containing compounds that either occur in living organisms or have characteristics that resemble compounds in living organisms are called organic chemicals. 
  • Minerals. A solid, natural substance in which atoms are arranged in an orderly pattern is a mineral. A single coherent sample of a mineral that grew to its present shape is a crystal, whereas an irregularly shaped sample, or a fragment derived from a once-larger crystal or cluster of crystals, is a grain. 
  • Glasses. A solid in which atoms are not arranged in an orderly pattern is called glass. 
  • Rocks. Aggregates of mineral crystals or grains, or masses of natural glass, are called rocks. Geologists recognize three main groups of rocks. (1) Igneous rocks develop when hot molten (liquid) rock cools and freezes solid. (2) Sedimentary rocks form from grains that break off pre-existing rock and become cemented together, or from minerals that precipitate out of a water solution. (3) Metamorphic rocks form when pre-existing rocks change in response to heat and pressure. 
  • Sediment. An accumulation of loose mineral grains (grains that have not stuck together) is called sediment. 
  • Metals. A solid composed of metal atoms (such as iron, aluminium, copper, and tin) is called a metal. An alloy is a mixture containing more than one type of metal atom. 
  • Melts. A melt forms when solid materials become hot and transform into liquid. Molten rock is a type of melt geologists distinguish between magma, which is molten rock beneath the Earth’s surface, and lava, molten rock that has flowed out onto the Earth’s surface. 
  • Volatiles. Materials that easily transform into gas at the relatively low temperatures found at the Earth’s surface are called volatiles. 
The most common minerals in the Earth contain silica (a compound of silicon and oxygen) mixed in varying proportions with other elements. These minerals are called silicate minerals. Not surprisingly, rocks composed of silicate minerals are silicate rocks. Geologists distinguish four classes of igneous silicate rocks based, in essence, on the proportion of silica to iron and magnesium. In order, from greatest to least proportion of silica to iron and magnesium, these classes are felsic (or silicic), intermediate, mafic, and ultramafic. As the proportion of silica in a rock increases, the density (mass per unit volume) decreases. Thus, felsic rocks are less dense than mafic rocks. For now, we introduce the four rock types whose names we need to know for our discussion of the Earth’s layers that follows. These are  
  1. granite, a felsic rock with large grains; 
  2. gabbro, a mafic rock with large grains; 
  3. basalt, a mafic rock with small grains; and 
  4. peridotite, an ultramafic rock with large grains.

Discovering the Earth’s Internal Layers 

People have speculated about what’s inside our planet since ancient times. What is the source of incandescent lavas that spew from volcanoes, of precious gems and metals found in mines, of sparkling mineral waters that bubble from springs, and of the mysterious forces that shake the ground and topple buildings? In ancient Greece and Rome, the subsurface was the underworld, Hades, home of the dead, a region of fire and sulphurous fumes. Perhaps this image was inspired by the molten rock and smoke emitted by the volcanoes of the Mediterranean region. In the 18th and 19th centuries, European writers thought the Earth’s interior resembled a sponge, containing open caverns variously filled with molten rock, water, or air. In fact, in the popular 1864 novel Journey to the Centre of the Earth, by the French author Jules Verne, three explorers hike through interconnected caverns down to the Earth’s centre. 
How can we explore the interior for real? We can’t dig or drill down very far. Indeed, the deepest mine penetrates only about 3.5 km beneath the surface of South Africa. And the deepest drill hole probes only 12 km below the surface of northern Russia compared with the 6,371 km radius of the Earth, this hole makes it less than 0.2% of the way to the centre and is nothing more than a pinprick. Our modern image of the Earth’s interior, one made up of distinct layers, is the end product of many discoveries made during the past 200 years.
The first clue that led away from Jules Verne’s sponge image came when researchers successfully measured the mass of the whole Earth, and from this information derived its average density. They found that the average density of our planet far exceeds the density of common rocks found on the surface. Thus, the interior of the Earth must contain denser material than its outermost layer and can’t possibly be full of holes. In fact, the mass of the Earth overall is so great that the planet must contain a large amount of metal. Since the Earth is close to being a sphere, the metal must be concentrated near the centre. Otherwise, centrifugal force due to the spin of the Earth on its axis would pull the equator out, and the  planet would become a disk. (To picture why, consider that when you swing a hammer, your hand feels more force if you hold the end of the light wooden shaft, rather than the heavy metal head.) Finally, researchers realized that, though molten rock occasionally oozes out of the interior at volcanoes, the interior must be mostly solid, because if it weren't, the land surface would rise and fall due to tidal forces much more than it does.

An early image of Earth’s internal layers. 
Eventually, researchers concluded that the Earth resembled a hard-boiled egg, in that it had three principal layers: a not-so-dense crust (like an eggshell, composed of rocks such as granite, basalt, and gabbro), a denser solid mantle in the middle (the “white,” composed of a then-unknown material), and a very dense core (the “yolk,” composed of an unknown metal) (figure above a, b). Clearly, many questions remained. How thick are the layers? Are the boundaries between layers sharp or gradational? And what exactly are the layers composed of?

Clues from the Study of Earthquakes: Refining the Image 

Faulting and earthquakes.
When rock within the outer portion of the Earth suddenly breaks and slips along a fracture called a fault, it generates shock waves (abrupt vibrations), called seismic waves, that travel through the surrounding rock outward from the break. Where these waves cause the surface of the Earth to vibrate, people feel an earthquake, an episode of ground shaking. You can simulate this process, at a small scale, when you break a stick between your hands and feel the snap with your hands (figure above a, b). 
In the late 19th century, geologists learned that earthquake energy could travel, in the form of waves, all the way through the Earth’s interior from one side to the other. Geologists immediately realized that the study of earthquake waves travelling through the Earth might provide a tool for exploring the Earth’s insides, much as ultrasound today helps doctors study a patient’s insides. Specifically, laboratory measurements demonstrated that earthquake waves travel at different velocities (speeds) through different materials. Thus, by detecting depths at which velocities suddenly change, geoscientists pinpointed the boundaries between layers and even recognized subtler boundaries within layers. For example, such studies led geoscientists to subdivide the mantle into the upper mantle and lower mantle, and subdivide the core into the inner core and outer core. 

Pressure and Temperature Inside the Earth 

In order to keep underground tunnels from collapsing under the pressure created by the weight of overlying rock, mining engineers must design sturdy support structures. It is no surprise that deeper tunnels require stronger supports: the downward push from the weight of overlying rock increases with depth, simply because the mass of the overlying rock layer increases with depth. In solid rock, the pressure at a depth of 1 km is about 300 atm. At the Earth’s centre, pressure probably reaches about 3,600,000 atm. 
Temperature also increases with depth in the Earth. Even on a cool winter’s day, miners who chisel away at gold veins exposed in tunnels 3.5 km below the surface swelter in temperatures of about 53°C (127°F). We refer to the rate of change in temperature with depth as the geothermal gradient. In the upper part of the crust, the geothermal gradient  averages between 20°C and 30°C per km. At greater depths, the rate decreases to 10°C per  km or less. Thus, 35 km below the surface of a continent, the temperature reaches 400°C to 700°C, and the mantle-core boundary is about 3,500°C. No one has ever directly measured the temperature at the Earth’s centre, but calculations suggest it may exceed 4,700°C, close to the Sun’s surface temperature of 5,500°C. 

السبت، 23 يناير 2016

We Are All Made of Stardust

We Are All Made of Stardust 

Where Do Elements Come From? 

Element factories in space.
Nebulae from which the first-generation stars formed consisted entirely of the lightest atoms, because only these atoms were generated by Big Bang nucleosynthesis. In contrast, the Universe of today contains 92 naturally occurring elements. Where did the other 87 elements come from? In other words, how did elements with larger atomic numbers (such as carbon, sulphur, silicon, iron, gold, and uranium), which are common on Earth, form? Physicists have shown that these elements form during the life cycle of stars, by the process of stellar nucleosynthesis. Because of stellar nucleosynthesis, we can consider stars to be “element factories,” constantly fashioning larger atoms out of smaller atoms. 
What happens to the atoms formed in stars? Some escape into space during the star’s lifetime, simply by moving fast enough to overcome the star’s gravitational pull. The stream of atoms emitted from a star during its lifetime is a stellar wind (figure above a). Some escape only when a star dies. A small or medium star (like our Sun) releases a large shell of gas as it dies, ballooning into a “red giant” during the process, whereas a large star blasts matter into space during a supernova explosion (figure above b). Most very heavy atoms (those with atomic numbers greater than that of iron) require even more violent circumstances to form than generally occurs within a star. In fact, most very heavy atoms form during a supernova explosion. Once ejected into space, atoms from stars and supernova explosions form new nebulae or mix back into existing nebulae.
When the first generation of stars died, they left a legacy of new, heavier elements that mixed with residual gas from the Big Bang. A second generation of stars and associated planets formed out of these compositionally more diverse nebulae. Second-generation stars lived and died, and contributed heavier elements to third-generation stars. Succeeding generations contain a greater proportion of heavier elements. Because not all stars live for the same duration of time, at any given moment the Universe contains many different generations of stars. Our Sun may be a third-, fourth-, or fifth-generation star. Thus, the mix of elements we find on Earth includes relicts of primordial gas from the Big Bang as well as the disgorged guts of dead stars. Think of it the elements that make up your body once resided inside a star!

The Nebular Theory for Forming  the Solar System 

In recent posts, we introduced scientific concepts of how stars form from nebulae. But we delayed our discussion of how the planets and other objects in our Solar System originated until we had discussed the production of heavier atoms such as carbon, silicon, iron, and uranium, because planets consist predominantly of these elements. Now that we’ve discussed stars as element factories, we return to the early  history of the Solar System and introduce the nebular theory, an explanation for the origin of planets, moons, asteroids, and comets. According to the nebular theory, these objects formed from the material in the flattened outer part of the disk, the material that did not become part of the star. This outer part is called the protoplanetary disk. 
What did the protoplanetary disk consist of? The disk from which our Solar System formed contained all 92 elements, some as isolated atoms, and some bonded to others in molecules. Geologists divide the material formed from these atoms and molecules into two classes. Volatile materials such as hydrogen, helium, methane, ammonia, water, and carbon dioxide are materials that can exist as gas at the Earth’s surface. In the pressure and temperature conditions of space, all volatile materials remain in a gaseous state closer to the Sun. But beyond a distance called the “frost line,” some volatiles condense into ice. (Note that we do not limit use of the word “ice” to water alone.) Refractory materials are those that melt only at high temperatures, and they condense to form solid soot-sized particles of “dust” in the coldness of space. As the proto-Sun began to form, the inner part of the disk became hotter, causing volatile elements to evaporate and drift to the outer portions of the disk. Thus, the inner part of the disk ended up consisting predominantly of refractory dust, whereas the outer portions accumulated large quantities of volatile materials and ice. As this was happening, the protoplanetary disk evolved into a series of concentric rings in response to gravity.

The grainy interior of this meteorite may resemble the texture of a small planetesimal.
How did the dusty, icy, and gassy rings transform into planets? Even before the proto-Sun ignited, the material of the surrounding rings began to clump and bind together, due to gravity and electrical attraction. First, soot-sized particles merged to form sand- to marble-sized grains that resembled “dust bunnies.” Then, these grains stuck together to form grainy basketball-sized blocks (figure above), which in turn collided. If the collision was slow, blocks stuck together or simply bounced apart. If the collision was fast, one or both of the blocks shattered, producing smaller fragments that recombined later. Eventually, enough blocks coalesced to form planetesimals, bodies whose diameter exceeded about 1 km. Because of their mass, the planetesimals exerted enough gravity to attract and pull in other objects that were nearby. Figuratively, planetesimals acted like vacuum cleaners, sucking in small pieces of dust and ice as well as smaller planetesimals that lay in their orbit, and in the process they grew progressively larger. Eventually, victors in the competition to attract mass grew into protoplanets, bodies approaching the size of today’s planets. Once a protoplanet succeeded in incorporating virtually all the debris within its orbit, it became a full-fledged planet. 
Early stages in Earth’s planet-forming process probably occurred very quickly some computer models suggest that it may have taken less than a million years to go from the dust and gas stage to the large planetesimal stage. Planets may have grown from planetesimals in 10 to 200 million years. In the inner orbits, where the protoplanetary disk consisted mostly of dust, small terrestrial planets composed of rock and metal formed. In the outer part of the Solar System, where significant amounts of ice existed, protoplanets latched on to vast amounts of ice and gas and evolved into the giant planets. Fragments of materials that were not incorporated in planets remain today as asteroids and comets.
When did the planets form? Geologists have found that special types of meteorites thought to be leftover planetesimals formed at 4.57 Ga, and thus consider that date to be the birth date of the Solar System. If this date is correct, it means that the Solar System formed about 9 billion years after the Big Bang, and thus is only about a third as old as the Universe. 

Differentiation of the Earth  and Formation of the Moon 

When planetesimals started to form, they had a fairly homogeneous distribution of material throughout, because the smaller pieces from which they formed all had much the same composition and collected together in no particular order. But large planetesimals did not stay homogeneous for long, because they began to heat up. The heat came primarily from three sources: the heat produced during collisions (similar to the phenomenon that happens when you bang on a nail with a hammer and they both get warm), the heat produced when matter is squeezed into a smaller volume, and the heat produced from the decay of radioactive elements. In bodies whose temperature rose sufficiently to cause internal melting, denser iron alloy separated out and sank to the centre of the body, whereas lighter rocky materials remained in a shell surrounding the centre. By this process, called differentiation, protoplanets and large planetesimals developed internal layering early in their history. As we will see later, the central ball of iron alloy constitutes the body’s core and the outer shell constitutes its mantle.
In the early days of the Solar System, planets continued to be bombarded by meteorites (solid objects, such as fragments of planetesimals, falling from space that land on a planet) even after the Sun had ignited and differentiation had occurred (see Meteors and Meteorites). Heavy bombardment in the early days of the Solar System pulverized the surfaces of planets and eventually left huge numbers of craters. Bombardment also contributed to heating the planets. 
Based on analysis and the dating of Moon rocks, most geologists have concluded that at about 4.53 Ga, a Mars-sized protoplanet slammed into the newborn Earth. In the process, the colliding body disintegrated and melted, along with a large part of the Earth’s mantle. A ring of debris formed around the remaining, now-molten Earth, and quickly coalesced to form the Moon. Not all moons in the Solar System necessarily formed in this manner. Some may have been independent protoplanets or comets that were captured by a larger planet’s gravity. 

Meteors and Meteorites 

During the early days of the Solar System,  the Earth collided with and incorporated countless planetesimals and smaller fragments of solid material lying in its path. Intense bombardment ceased about  3.9 Ga, but even today collisions with space objects continue, and over 1,000 tons of material (rock, metal, dust, and ice) fall to Earth, on average, every year. The vast majority of this material consists of fragments derived from comets and asteroids sent careening into the path of the Earth after billiard-ball-like collisions with each other out in space, or because of the gravitational pull of a passing planet deflected their orbit. Some of the material, however, consists of chips of the Moon or Mars, ejected into space when large objects collided with those bodies. 

Meteors and meteorites.
Astronomers refer to any object from space that enters the Earth’s atmosphere as a meteoroid. Meteoroids move at speeds of 20 to 75 km/s (over 45,000 mph), so fast that when they reach an altitude of about 150 km, friction with the atmosphere causes them to heat up and vaporize, leaving a streak of bright, glowing gas. The glowing streak, an atmospheric  phenomenon, is a meteor (also known colloquially, though incorrectly, as a “falling star”) (figure above a). Most visible meteors completely vaporize by an altitude of about 30 km. But dust-sized ones may slow down sufficiently to float to Earth, and larger ones (fist-sized or bigger) can survive the heat of entry to reach the surface of the planet. In some cases, meteoroids explode in brilliant fireballs. 
Objects that strike the Earth are called meteorites. Although almost all meteorites are small and have not caused notable damage on Earth during human history, a very few have smashed through houses, dented cars, and bruised people. During the longer term of Earth history, however, there have been some catastrophic collisions that left huge craters (figure above b). 
Most meteorites are asteroidal or planetary fragments, for icy material is too fragile to survive the fall. Researchers recognize three basic classes of meteorites: iron (made of iron-nickel alloy), stony (made of rock), and stony iron (rock embedded in a matrix of metal). Of all known meteorites, about 93% are stony and 6% are iron (figure above c). Researchers have concluded that some meteors (a special subcategory of stony meteorites called carbonaceous chondrites, because they contain carbon and small spherical nodules called chondrules) are asteroids derived from planetesimals that never underwent differentiation into a core and mantle. Other stony meteorites and all iron meteorites are asteroids derived from planetesimals that had differentiated into a metallic core and a rocky mantle early in Solar System history but later shattered into fragments during collisions with other planetesimals. Most meteorites appear to be about 4.54 Ga, but carbonaceous chondrites are as old as 4.57 Ga and are the oldest solar system materials ever measured.
Since meteorites represent fragments of undifferentiated and differentiated planetesimals, geologists consider the average composition of meteorites to be representative of the average composition of the whole Earth. In other words, the estimates that geologists use for the proportions of different elements in the Earth are based largely on studying meteorites. Stony meteorites are probably similar in composition to the mantle, and iron meteorites are probably similar in composition to the core.

Why Are Planets Round? 

Small planetesimals were jagged or irregular in shape, and asteroids today have irregular shapes. Planets, on the other hand, are more or less spherical. Why? Simply put, when a protoplanet gets big enough, gravity can change its shape. To picture how, imagine a block of cheese sitting outside on a hot summer day. As the cheese gets softer and softer, gravity causes it to spread out in a pancake-like blob. This model shows that gravitational force alone can cause material to change shape if the material is soft enough. Now let’s apply this model to planetary growth. 
The rock composing a small planetesimal is cool and strong enough so that the force of gravity is not sufficient to cause the rock to flow. But once a planetesimal grows beyond a certain critical size (about 1,000 km in diameter), its interior becomes warm and soft enough to flow in response to gravity. As a consequence, protrusions are pulled inward toward the centre, and the planetesimal re-forms into a special shape that permits the force of gravity to be nearly the same at all points on its surface. This special shape is spherical because in a sphere mass is evenly distributed around the centre.
Credits: Stephen Marshak (Essentials of Geology)

الأربعاء، 20 يناير 2016

Universe formation

Universe formation

We stand on a planet, in orbit around a star, speeding through space on the arm of a galaxy. Beyond our galaxy lie hundreds of billions of other galaxies. Where did all this “stuff” the matter of the Universe come from, and when did it first form? For most of human history, a scientific solution to these questions seemed intractable. But in the 1920s, unexpected observations about the nature of light from distant galaxies set astronomers on a path of discovery that ultimately led to a model of Universe formation known as the Big Bang theory. To explain these observations, we must first introduce an important phenomenon called the Doppler effect. We then show how this understanding leads to the recognition that the Universe is expanding, and finally, to the conclusion that this expansion began during the Big Bang, 13.7 billion years ago.

Waves and the Doppler Effect 

When a train whistle screams, the sound you hear moved through the air from the whistle to your ear in the form of sound waves. Waves are disturbances that transmit energy from one point to another in the form of periodic motions. As each sound wave passes, air alternately compresses, then expands. We refer to the distance between successive waves as the wavelength, and the number of waves that pass a point in a given time interval as the frequency. If the wavelength decreases, more waves pass a point in a given time interval, so the frequency increases. The pitch of a sound, meaning its note on the musical scale, depends on the frequency of the sound waves. 

Manifestations of the Doppler effect for sound and for light.
Imagine that you are standing on a station platform while a train moves toward you. The train whistle’s sound gets louder as the train approaches, but its pitch remains the same. The instant the train passes, the pitch abruptly changes it sounds like a lower note in the musical scale. Why? When the train moves toward you, the sound has a higher frequency (the waves are closer together so the  wavelength is smaller) because the sound source, the whistle, has moved slightly closer to you between the instant that it emits one wave and the instant that it emits the next  (figure above a, b). When the train moves away from you, the sound has a lower frequency (the waves are farther apart), because the whistle has moved slightly farther from you between the instant it emits one wave and the instant it emits the next. An Austrian physicist, C. J. Doppler (1803–1853), first interpreted this phenomenon, and thus the change in frequency that happens when a wave source moves is now known as the Doppler effect. 
Light energy also moves in the form of waves. We can represent light waves symbolically by a periodic succession of crests and troughs (figure above c). Visible light comes in many colours the colours of the rainbow. The colour you see depends on the frequency of the light waves, just as the pitch of a sound you hear depends on the frequency of sound waves. Red light has a longer wavelength (lower frequency) than does blue light. The Doppler effect also applies to light but can be noticed only if the light source moves very fast, at least a few percent of the speed of light. If a light source moves away from you, the light you see becomes redder, as the light shifts to longer wavelength or lower frequency. If the source moves toward you, the light you see becomes bluer, as the light shifts to higher frequency. We call these changes the red shift and the blue shift, respectively.

Does the Size of the Universe Change? 

In the 1920s, astronomers such as Edwin Hubble, after whom the Hubble Space Telescope was named, braved many a frosty night beneath the open dome of a mountaintop observatory in order to aim telescopes into deep space. These researchers were searching for distant galaxies. At first, they documented only the location and shape of newly discovered galaxies, but eventually they also began to study the wavelength of light produced by the distant galaxies. The results yielded a surprise that would forever change humanity’s perception of the Universe. To their amazement, the astronomers found that the light of distant galaxies display a red shift relative to the light of a nearby star (figure above d). 
Hubble pondered this mystery and, around 1929, attributed the red shift to the Doppler effect, and concluded that the distant galaxies must be moving away from Earth at an immense velocity. At the time, astronomers thought the Universe had a fixed size, so Hubble initially assumed that if some galaxies were moving away from Earth, others must be moving toward Earth. But this was not the case. On further examination, Hubble concluded that the light from all distant galaxies, regardless of their direction from Earth, exhibits a red shift. In other words, all distant galaxies are moving rapidly away from us. 

 The concept of the expanding Universe and the Big Bang.
How can all galaxies be moving away from us, regardless of which direction we look? Hubble puzzled over this question and finally recognized the solution: the whole Universe must be expanding! To picture the expanding Universe, imagine a ball of bread dough with raisins scattered throughout. As the dough bakes and expands into a loaf, each raisin moves away from its neighbours, in every direction;  figure above a. This idea came to be known as the expanding Universe theory.

The Big Bang 

Hubble’s ideas marked a revolution in cosmological thinking. Now we picture the Universe as an expanding bubble, in which galaxies race away from each other at incredible speeds. This image immediately triggers the key question of cosmology: did the expansion begin at some specific time in the past? If it did, then that instant would mark the physical beginning of the Universe. 
Most astronomers have concluded that expansion did indeed begin at a specific time, with a cataclysmic explosion called the Big Bang. According to the Big Bang theory, all matter and energy everything that now constitutes the  Universe was initially packed into an infinitesimally small point. The point “exploded” and the Universe began, according to current estimates, 13.7 (+-1%) billion years ago.
Of course, no one was present at the instant of the Big Bang, so no one actually saw it happen. But by combining clever calculations with careful observations, researchers have developed a consistent model of how the Universe evolved, beginning an instant after the explosion (figure above b). According to this model of the Big Bang, profound change happened at a fast and furious rate at the outset. During the first instant of existence, the Universe was so small, so dense, and so hot that it consisted entirely of energy atoms, or even the smallest subatomic particles that make up atoms, could not even exist. (See nature of matter for a review of atomic structure.) Within a few seconds, however, hydrogen atoms could begin to form. And by the time the Universe reached an age of 3 minutes, when its temperature had fallen below 1 billion degrees, and its diameter had grown to about 53 million km (35 million miles), hydrogen atoms could fuse together to form helium atoms. Formation of new nuclei in the first few minutes of time is called Big Bang nucleosynthesis because it happened before any stars existed. This process could produce only light atoms, meaning ones containing a small number of protons (an atomic number less than 5), and it happened very rapidly. In fact, virtually all of the new atomic nuclei that would form by Big Bang n ucleosynthesis existed by the end of the first 5 minutes. 
Eventually, the Universe became cool enough for chemical bonds to bind atoms of certain elements together in molecules. Most notably, two hydrogen atoms could join to form molecules of H2. As the Universe continued to expand and cool further, atoms and molecules slowed down and accumulated into patchy clouds called nebulae. The earliest nebulae of the Universe consisted almost entirely of hydrogen (74%, by volume) and helium (24%) gas.

The Nature of Matter

The nature of atoms and nuclear reactions.
What does matter consist of? A Greek philosopher named Democritus (ca. 460–370 B.C.E.) argued that if you kept dividing matter into progressively smaller pieces, you would eventually end up with nothing; but since it’s not possible to make something out of nothing, there must be a smallest piece of matter that can’t be subdivided further. He proposed the name “atom” for these smallest pieces, based on the Greek word atomos, which means indivisible. Our modern understanding of matter developed in the 17th century, when chemists recognized that certain substances (such as hydrogen and oxygen) cannot break down into other substances, whereas others (such as water and salt) can break down. The former came to be known as elements, and the latter came to be known as compounds. John Dalton (1766–1844) adopted the word atom for the smallest piece of an element that has the property of the element; the smallest piece of a compound that has the properties of the compound is a molecule. Separate atoms are held together to form molecules by chemical bonds, which we discuss more fully later in the book. As an example, chemical bonds hold two hydrogen atoms to form an H2 molecule. Chemists in the 17th and 18th centuries identified 92 naturally occurring elements on Earth; modern physicists have created more than a dozen new ones. Each element has a name and a symbol (e.g., N = nitrogen; H = hydrogen; Fe = iron; Ag = silver). Atoms are so small that over five trillion (5,000,000,000,000) can fit on the head of a pin. Nevertheless, in 1910, Ernest Rutherford, a British physicist, proved that, contrary to the view of Democritus, atoms actually can be divided into smaller pieces. Most of the mass in an atom clusters in a dense ball, called the nucleus, at the atom’s centre. The nucleus contains two types of subatomic particles: neutrons, which have a neutral electrical charge, and protons, which have a positive charge. A cloud of electrons surrounds the nucleus (figure above a); an electron has a negative charge and contains only 1/1,836 as much mass as a proton. (“Charge,” simplistically, refers to the way in which a particle responds to a magnet or an electric current.) Roughly speaking, the diameter of an electron cloud is 10,000 times greater than that of the nucleus, yet the cloud contains only 0.05% of an atom’s mass thus, atoms are mostly empty space! We distinguish atoms of different elements from one another by their atomic number, the number of protons in their nucleus. Smaller atoms have smaller atomic numbers, and larger ones have larger atomic numbers. The lightest atom, hydrogen, has an atomic number of 1, and the heaviest naturally occurring atom, uranium, has an atomic number of 92. Except for hydrogen nuclei, all nuclei also contain neutrons. In smaller atoms, the number of neutrons roughly equals the number of protons, but in larger atoms the number of neutrons exceeds the number of protons. The atomic mass of an atom is roughly the sum of the number of neutrons and the number of  protons. For example, an oxygen nucleus contains 8 protons and 8 neutrons, and thus has an atomic mass of 16. In 1869, a Russian chemist named Dmitri Mendelév (1834–1907) recognized that groups of elements share similar characteristics, and he organized the elements into a chart that we now call the periodic table of the elements. With modern understanding of the periodic table, it became clear that the ordering of the elements reflects their atomic number and the stream of the electron cloud. Nuclear bonds serve as the “glue” that holds together subatomic particles in a nucleus. Atoms can change only during nuclear reactions, when nuclear bonds break or form. Physicists recognize several types of nuclear reactions. For example, during “radioactive decay” reactions, a nucleus either emits a subatomic particle or undergoes fission. As a result of fission, a large nucleus breaks apart to form two smaller atoms (figure above b). Radioactive decay transforms an atom of one element into an atom of another and produces energy. For example, fission reactions provide the energy of atomic bombs and nuclear power plants. Atoms that spontaneously undergo the process are known as radioactive elements. During fusion, smaller atoms collide and stick together to form a larger atom. For example, successive fusion reactions produce a helium atom out of four hydrogen atoms. Fusion reactions power the Sun and occur during the explosion of a hydrogen bomb (figure above c).

Birth of the First Stars 

When the Universe reached its 200 millionth birthday, it contained immense, slowly swirling, dark nebulae separated by vast voids of empty space. The Universe could not remain this way forever, though, because of the invisible but persistent pull of gravity. Eventually, gravity began to remold the Universe pervasively and permanently. 
All matter exerts gravitational pull a type of force on its surroundings, and as Isaac Newton first pointed out, the amount of pull depends on the amount of mass; the larger the mass, the greater its pull. Somewhere in the young Universe, the gravitational pull of an initially more massive region of a nebula began to suck in surrounding gas and, in a grand example of the rich getting richer, grew in mass and, therefore, density. As this denser region attracted progressively more gas, the gas compacted into a smaller region, and the initial swirling movement of gas transformed into a rotation around an axis. As gas continued to move inward, cramming into a progressively smaller volume, the rotation rate became faster and faster. (A similar phenomenon happens when a spinning ice skater pulls her arms inward.) Because of its increased rotation, the nebula evolved into a disk shape (see Geology at a Glance, pp. 22–23). As more and more matter rained down onto the disk, it continued to grow, until eventually, gravity collapsed the inner portion of the disk into a dense ball. As the gas squeezed into a smaller and smaller space, its temperature increased dramatically. Eventually, the central ball of the disk became hot enough to glow, and at this point it became a protostar. The remaining mass of the disk, as we will see, eventually clumped into smaller spheres, the planets. 
A protostar continues to grow, by pulling in successively more mass, until its core becomes extremely dense and its temperature reaches about 10 million degrees. Under such conditions, hydrogen nuclei slam together so forcefully that they join or “fuse,” in a series of steps, to form helium nuclei (see the nature of matter).  Such fusion reactions produce huge amounts of energy, and the mass becomes a fearsome furnace. When the first nuclear fusion reactions began in the first protostar, the body “ignited” and the first true star formed. When this happened, perhaps 800 million years after the Big Bang, the first starlight pierced the newborn Universe. This process would soon happen again and again, and many first-generation stars came into existence. 
First-generation stars tended to be very massive, perhaps 100 times the mass of the Sun. Astronomers have shown that the larger the star, the hotter it burns and the faster it runs out of fuel and dies. A huge star may survive only a few million years to a few tens of millions of years before it becomes a supernova, a giant explosion that blasts much of the star’s matter back into space. Thus, not long after the first generation of stars formed, the Universe began to be peppered with the first generation of supernovas.
Credit: Stephen Marshak (Essentials of Geology)

السبت، 16 يناير 2016

An Image of Our Universe

An Image of Our Universe

What Is the Structure of the Universe? 

Contrasting views of the universe drawn by artist hundreds of years ago.
Think about the mysterious spectacle of a clear night sky. What objects are up there? How big are they? How far away are they? How do they move? How are they arranged? In addressing such questions, ancient philosophers first distinguished between stars (points of light whose locations relative to each other are fixed) and planets (tiny spots of light that move relative to the backdrop of stars). Over the centuries, two schools of thought developed concerning how to explain the configuration of stars and planets, and their relationships to the Earth, Sun, and Moon. The first school advocated a geocentric model (figure above a), in which the Earth sat without moving at the centre of the Universe, while the Moon and the planets whirled around it within a revolving globe of stars. The second school advocated a heliocentric model (figure above b), in which the Sun lay at the centre of the Universe, with the Earth and other planets orbiting around it.
The geocentric image eventually gained the most followers, due to the influence of an Egyptian mathematician,  Ptolemy (100–170 C.E.), for he developed equations that appeared to predict the wanderings of the planets in the context of the geocentric model. During the Middle Ages (ca. 476–1400 C.E.), church leaders in Europe adopted Ptolemy’s geocentric model as dogma, because it justified the comforting thought that humanity’s home occupies the most important place in the Universe. Anyone who disagreed with this view risked charges of heresy. 
Then came the Renaissance. In 15th-century Europe, bold thinkers spawned a new age of exploration and scientific discovery. Thanks to the efforts of Nicolaus Copernicus (1473–1543) and Galileo Galilei (1564–1642), people gradually came to realize that the Earth and planets did indeed orbit the Sun and could not be at the centre of the Universe. And when Isaac Newton (1643–1727) explained gravity, the attractive force that one object exerts on another, it finally became possible to understand why these objects follow the orbits that they do. 
In the centuries following Newton, scientists gradually adopted modern terminology for discussing the Universe. In this language, the Universe contains two related entities: matter and energy. Matter is the substance of the Universe it takes up space and you can feel it. We refer to the amount of matter in an object as its mass, so an object with greater mass contains more matter. Density refers to the amount of mass occupying a given volume of space. The mass of an object determines its weight, the force that acts on an object due to gravity. 
An object always has the same mass, but its weight varies depending on where it is. For example, on the Moon, you weigh much less than on the Earth. The matter in the Universe does not sit still. Components vibrate and spin, they move  from one place to another, they pull on or push against each other, and they break apart or combine. In a general sense, we consider such changes to be kinds of “work.” Physicists refer to the ability to do work as energy. One piece of matter can do work directly on another by striking it. Heat, light, magnetism, and gravity all provide energy that can cause change at a distance. 

A galaxy may contain about 300 billion stars.
As the understanding of matter and energy improved, and telescopes became refined so that astronomers could see and measure features progressively farther into space, the interpretation of stars evolved. Though it looks like a point of light, a star is actually an immense ball of incandescent gas that emits intense heat and light. Stars are not randomly scattered through the Universe; gravity holds them together in immense groups called galaxies. The Sun and over 300 billion stars together form the Milky Way galaxy. More than 100 billion galaxies constitute the visible Universe (figure above a). 
From our vantage point on Earth, the Milky Way looks like a hazy band (figure above b), but if we could view the Milky Way from a great distance, it would look like a flattened spiral with great curving arms slowly swirling around a glowing, disk-like centre (figure above c). Presently, our Sun lies near the outer edge of one of these arms and rotates around the centre of the galaxy about once every 250 million years. So, we hurtle through space, relative to an observer standing outside the galaxy, at about 200 km per second. 
Clearly, human understanding of Earth’s place in the Universe has evolved radically over the past few centuries. Neither the Earth, nor the Sun, nor even the Milky Way occupy the centre of the Universe and everything is in motion. 

The Nature of Our Solar System 

Eventually, astronomical study demonstrated that our Sun is a rather ordinary, medium-sized star. It looks like a sphere, instead of a point of light, because it is much closer to the Earth than are the stars. The Sun is “only” 150 million km (93  million miles) from the Earth. Stars are so far away that we measure their distance in light years, where 1 light year is the distance travelled by light in one year, about 10 trillion km, or 6 trillion miles the nearest star beyond the Sun is over 4 light years away. How can we picture distances? If we imagined that the Sun were the size of a golf ball (about 4.3 cm), then the Earth would be a grain of sand about one meter away, and the nearest star would be 270 km (168 miles) away. (Note that the distance between stars is tiny by galactic standards, the Milky Way galaxy is 120,000 light years across!)

 The relative sizes and positions of planets in the Solar System.
Our Sun is not alone as it journeys through the heavens. Its gravitational pull holds on to many other objects which, together with the Sun, comprise the Solar System (figure above a, b). The Sun accounts for 99.8% of the mass in the Solar System. The remaining 0.2% includes a great variety of objects, the largest of which are planets. Astronomers define a planet as an object that orbits a star, is roughly spherical, and has “cleared its neighbourhood of other objects.” The last phrase in this definition sounds a bit strange at first, but merely implies that a planet’s gravity has pulled in all particles of matter in its orbit.
According to this definition, which was formalized in 2005, our Solar System includes eight planets Mercury, Venus, Earth, Mars, Jupiter, Saturn, Uranus, and Neptune. In 1930, astronomers discovered Pluto, a 2,390-km-diameter sphere of ice, whose orbit generally lies outside that of Neptune’s. Until 2005, astronomers considered Pluto to be a planet. But since it does not fit the modern definition, it has been dropped from the roster. Our Solar System is not alone in hosting planets; in recent years, astronomers have found planets orbiting stars in many other systems. As of 2012, over 760 of these “exoplanets” have been found. 
Planets in our Solar System differ radically from one another both in size and composition. The inner planets (Mercury, Venus, Earth, and Mars), the ones closer to the Sun, are relatively small. Astronomers commonly refer to these as terrestrial planets because, like Earth, they consist of a shell of rock surrounding a ball of metallic iron alloy. The outer planets (Jupiter, Saturn, Uranus, and Neptune) are known as the giant planets, or Jovian planets. The adjective giant certainly seems appropriate, for these planets are huge Jupiter, for example, has a mass 318 times larger than that of Earth and accounts for about 71% of the non-solar mass in the Solar System. The overall composition of the giant planets is very different from that of the terrestrial planets. Specifically, most of the mass of Neptune and Uranus contain solid forms of water, ammonia, and methane, so these planets are known as the ice giants. Most of the mass of Jupiter and Saturn consists of hydrogen and helium gas or liquefied gas, so these planets are known as the gas giants.
In addition to the planets, the Solar System contains a great many smaller objects. Of these, the largest are moons. A moon is a sizeable body locked in orbit around a planet. All but two planets (Mercury and Venus) have moons in varying numbers Earth has one, Mars has two, and Jupiter has at least 63. Some moons, such as Earth’s Moon, are large and spherical, but most are small and have irregular shapes. In addition to moons, millions of asteroids (chunks of rock and/or metal) comprise a belt between the orbits of Mars and Jupiter. Asteroids range in size from less than a centimetre to about 930 km in diameter. And about a trillion bodies of ice lie in belts or clouds beyond the orbit of Neptune. Most of these icy objects are tiny, but a few (including Pluto) have diameters of over 2,000 km and may be thought of as “dwarf planets.” The gravitational pull of the main planets has sent some of the icy objects on paths that take them into the inner part of the Solar System, where they begin to evaporate and form long tails of gas we call such objects comets.
Credits: Stephen Marshak

الخميس، 7 يناير 2016

Desert Landscapes and Life

Desert Landscapes and Life 

The popular media commonly portray deserts as endless vistas of sand, punctuated by the occasional palm-studded oasis. In reality, not all desert landscapes are buried by sand. Some deserts  are vast, rocky plains; others sport a stubble of cacti and other hardy desert plants; and still others display intricate rock formations that look like medieval castles. Explorers of the Sahara, for example, traditionally distinguished among hamada (barren, rocky highlands), reg (vast, stony plains), and erg (sand seas in which large dunes form). 
In this post, we’ll see how the erosional and  depositional processes described above lead to the formation of such contrasting landscapes.

Rocky Cliffs and Mesas 

In hilly desert regions, the lack of soil exposes rocky ridges and cliffs. As noted earlier, cliffs erode when rocks split away along joints. When this happens, the cliff face steps back into the land but retains roughly the same form. The process, commonly referred to as cliff retreat, or scarp retreat, occurs in fits and starts. A cliff may remain unchanged for decades or centuries, and then suddenly a block of rock falls off and crumbles into rubble at the foot of the cliff. Cliffs exposing alternating layers of strata with contrasting strength develop a step-like shape; strong layers (sandstone or limestone) become vertical cliffs, and weak layers (shale) become rubble-covered slopes. 

Mesas and buttes form in deserts as cliffs retreat over time.
With continued erosion and cliff retreat, a plateau of rock slowly evolves into a cluster of isolated hills, ridges, or columns. Flat-lying strata or flat-lying layers of volcanic rocks erode to make flat-topped hills, which go by different names depending on their size. Large examples (with a top surface area of several square km) are mesas, from the Spanish word for table. Medium-sized examples are buttes (figure above a, b), and small examples, whose height greatly exceeds their top surface area, are chimneys (figure above c). Natural arches form when erosion along joints leaves narrow walls of rock. When the lower part of the wall erodes while the upper part remains, an arch results. 

 The formation of cuestas and inselbergs, due to erosion and deposition in deserts.
In places where bedding dips at an angle, an asymmetric ridge called a cuesta develops. A joint-controlled cliff forms the steep front side of a cuesta, and the tilted top surface of a resistant bed forms the gradual slope on the backside (figure above a). If the bedding dip is steep to near vertical, a narrow symmetrical ridge, called a hogback, forms. 
After a long period of erosion, all that may remain of a once broad region of uplifted land is a relatively small island of rock, surrounded by alluvium-filled basins. Geologists refer to such islands of rock by the German word inselberg (island mountain; figure above b). Depending on the rock type or the orientation of stratification in the rock, and on rates of erosion, inselbergs may be sharp-crested, plateau-like, or loaf-shaped (steep sides and a rounded crest). Inselbergs with a loaf geometry, as exemplified by Uluru (Ayers Rock) in central Australia (figure above c), are also known as bornhardts.

Desert Pavement 

Desert pavement, and a hypothesis for how it forms by building up a soil from below.
In many locations, the desert surface resembles a tile mosaic in that it consists of separate stones that fit together tightly, forming a fairly smooth surface layer above a soil composed of silt and clay. Such natural mosaics constitute desert pavement (figure above a, b). Typically, desert varnish coats the top surfaces of the stones forming desert pavement. Recently, researchers have suggested that pavements form when windblown dust slowly sifts down onto the stones and then washes down between the stones. In this model, the pavement is “born at the surface,” meaning that the stones forming the pavement were never buried, but have been progressively lifted up as soil collects and builds up beneath (figure above c). Over time, the rocks at the surface crack, perhaps due to differential heating by the desert sun. Sheetwash, during downpours, may wash away fine sediment between fragments, and when soils dry and shrink between storms, the clasts settle together, locking into a stable, jigsaw-like arrangement.

Stony Plains and Pediments 

The coarse sediment eroded from desert mountains and ridges washes into adjacent lowlands and builds out to form gently sloping alluvial fans. The surfaces of these gravelly piles are strewn with pebbles, cobbles, and boulders, and are dissected by dry washes (wadis or arroyos). Portions of these stony plains may evolve into desert pavements. 
When travelers began trudging through the desert of the southwestern United States during the 19th century, they found that in many locations the wheels of their wagons were rolling over flat or gently sloping bedrock surfaces. These bedrock surfaces extended outward like ramps from the steep cliffs of a mountain range on one side, to the alluvium-filled valleys on the other (see 2nd figure). Geologists now refer to such surfaces as pediments. Pediments develop when sheetwash during floods carries sediment away from the mountain front, during mountain-front retreat. The moving sediment grinds away the bedrock that it tumbles over.

Seas of Sand: The Nature of Dunes 

The types of sand dunes and the cross beds within them.
A sand dune is a pile of sand deposited by a moving current. Dunes in deserts may start to form where sand becomes trapped on the windward side of an obstacle, such as a rock or a shrub. Gradually, the sand builds downwind into the lee of the obstacle. Dunes display a variety of shapes and sizes, depending on the character of the wind and the sand supply (figure above a). Where the sand is relatively scarce and the wind blows steadily in one direction, beautiful crescents called barchan dunes develop, with the tips of the crescents pointing downwind. If the wind shifts direction frequently, a group of crescents pointing in different directions overlap one another, creating a constantly changing star dune. Where enough sand accumulates to bury the ground surface completely, and only moderate winds blow, sand piles into simple, wave-like shapes called transverse dunes. The crests of transverse dunes lie perpendicular to the wind direction. Strong winds may break through transverse dunes and change them into parabolic dunes whose ends point in the upwind direction. Finally, if there is abundant sand and a strong, steady wind, the sand streams into longitudinal dunes, whose axis lies parallel to the wind direction.
In a sand dune, sand saltates up the windward side of the dune, blows over the crest of the dune, and then settles on the steeper, lee face of the dune. The slope of this face attains the angle of repose, the slope angle of a free-standing pile of sand. As sand collects on this surface, it eventually becomes unstable and slides down the slope, so geologists refer to the lee side of a dune as the slip face. As more and more sand accumulates on the slip face, the crest of the dune migrates downwind, and former slip faces become preserved inside the dune. In cross section, these slip faces appear as cross beds (figure above b, c).
Figures credited to Stephen Marshak.

Interested for our works and services?
Get more of our update !