The Haber process

?
Intros
Lessons
  1. Equilibrium in real life chemistry.
  2. The Haber process: introduction.
  3. Changing conditions in the Haber process.
  4. How do we maximise production?
Topic Notes
?

Introduction to the Haber Process

Welcome to our exploration of the Haber process, a groundbreaking chemical reaction that revolutionized fertilizer production and transformed global agriculture. This ingenious method, developed by Fritz Haber and Carl Bosch in the early 20th century, enables the efficient synthesis of ammonia from nitrogen and hydrogen. The Haber process has become the cornerstone of modern fertilizer manufacturing, playing a crucial role in feeding the world's growing population. By dramatically increasing crop yields, it has had an immense impact on global food supply and security. Our introductory video will guide you through the basics of this process, helping you understand its significance in chemistry and its real-world applications. As we delve into the intricacies of the Haber process, you'll gain valuable insights into how science and technology can address global challenges. Let's embark on this fascinating journey together!

The Haber Process Equation and Reaction

The Haber process, also known as the Haber-Bosch process, is a crucial chemical reaction in modern industry. This process is fundamental to the production of ammonia, which is essential for fertilizers and various other applications. The Haber process equation can be written as:

N + 3H 2NH

This equation represents the reaction between nitrogen (N) and hydrogen (H) to produce ammonia (NH). The double arrow () indicates that this is a reversible reaction, meaning it can proceed in both directions under certain conditions.

The reactants in the Haber process are nitrogen and hydrogen. Nitrogen is typically obtained from the air, which is approximately 78% nitrogen by volume. Hydrogen, on the other hand, is usually produced through the steam reforming of natural gas or other hydrocarbon sources. The product of this reaction is ammonia, a compound of nitrogen and hydrogen with the formula NH.

One of the key characteristics of the Haber process is that it is an exothermic reaction. This means that heat is released when the reaction occurs. The exothermic nature of the reaction has important implications for its industrial application. As the reaction proceeds and heat is released, the equilibrium shifts towards the reactants, potentially reducing the yield of ammonia. To counteract this and maximize ammonia production, the process is carefully controlled using specific temperatures, pressures, and catalysts.

The development of the Haber process is credited to Fritz Haber and Karl Bosch in the early 20th century. Fritz Haber, a German chemist, first developed the process for synthesizing ammonia from its elements in 1909. He demonstrated that under high pressure and temperature, and with the use of an iron catalyst, it was possible to combine nitrogen and hydrogen to produce ammonia efficiently.

Karl Bosch, a German chemical engineer, then worked on scaling up Haber's laboratory process for industrial production. Bosch overcame significant engineering challenges, including the development of high-pressure equipment capable of withstanding the extreme conditions required for the reaction. The first Haber-Bosch plant began operation in 1913, marking a significant milestone in industrial chemistry.

The Haber process revolutionized agriculture by making the mass production of nitrogen-based fertilizers possible. This led to substantial increases in crop yields worldwide, supporting population growth and changing the face of modern agriculture. Today, the Haber-Bosch process is responsible for producing over 150 million tons of ammonia annually, consuming about 1-2% of the world's energy supply.

In industrial settings, the Haber process typically operates at temperatures between 400-450°C (752-842°F) and pressures of 150-300 atmospheres. An iron catalyst is used to increase the rate of reaction. The high pressure is necessary to shift the equilibrium towards the product side, increasing the yield of ammonia. However, the high temperature, which is needed to increase the reaction rate, tends to favor the reverse reaction. This balance between temperature and pressure is crucial for optimizing ammonia production.

Understanding the Haber process equation and reaction is essential for students of chemistry and chemical engineering. It demonstrates key concepts in chemical equilibrium, reaction kinetics, and industrial process design. The process also serves as an excellent example of how scientific discoveries can have profound impacts on society, economy, and the environment.

Conditions and Principles of the Haber Process

The Haber process, a cornerstone of modern industrial chemistry, operates under carefully controlled conditions to maximize the production of ammonia from nitrogen and hydrogen gases. Understanding these optimal conditions and the principles behind them is crucial for students and professionals alike in the field of chemical engineering.

Temperature plays a pivotal role in the Haber process. The reaction is typically carried out at temperatures ranging from 400°C to 450°C. This might seem counterintuitive at first, as the reaction is exothermic, meaning it releases heat. According to Le Chatelier's principle, which states that a system at equilibrium will adjust to counteract any change imposed upon it, increasing temperature should favor the reverse reaction. However, the high temperature is necessary to overcome the activation energy barrier and increase the rate of reaction. It's like trying to bake a cake - you need the oven hot enough to start the chemical reactions, even if a cooler temperature might theoretically yield a better result in the long run.

Pressure is another critical factor in the Haber process. The reaction is typically conducted at pressures between 150-250 atmospheres. This high pressure is essential because the forward reaction produces fewer gas molecules than the reactants (3 volumes of gas react to form 2 volumes). Again, applying Le Chatelier's principle, we can understand that increasing pressure will favor the reaction that produces fewer gas molecules, thus pushing the equilibrium towards ammonia production. Think of it as squeezing a balloon filled with nitrogen and hydrogen - the gases are forced to combine into a more compact form (ammonia) under the pressure.

The use of catalysts is the third key component in optimizing the Haber process. Iron-based catalysts, often promoted with small amounts of other elements like potassium, are commonly used. These catalysts don't change the position of equilibrium but dramatically increase the rate at which equilibrium is reached by lowering the activation energy barrier of the reaction. It's akin to a matchmaker in a crowded room - the catalyst brings the reactant molecules together more efficiently, speeding up their interaction without changing the final outcome.

The interplay of these conditions - high temperature, high pressure, and the presence of a catalyst - creates a delicate balance in the Haber process. The high temperature increases reaction rate but decreases yield, while high pressure increases yield but comes with significant engineering challenges and energy costs. The catalyst helps mitigate some of these trade-offs by allowing the reaction to proceed more quickly at lower temperatures than would otherwise be possible.

Le Chatelier's principle is the guiding concept behind understanding how these conditions affect the equilibrium of the Haber process. For instance, if we were to suddenly increase the concentration of nitrogen in the reaction chamber, the equilibrium would shift to use up some of this excess, producing more ammonia. Similarly, removing ammonia as it's produced shifts the equilibrium to produce more, following the principle's prediction that the system will act to counteract the change.

To visualize this, imagine a see-saw with nitrogen and hydrogen on one side and ammonia on the other. The conditions of the Haber process are like carefully adjusting the fulcrum and applying pressure to different points to keep the see-saw tipped towards the ammonia side, all while ensuring it keeps moving back and forth to maintain the reaction.

The yield of the reaction is directly impacted by these conditions. While the theoretical yield of ammonia under ideal conditions could approach 100%, practical limitations mean that a single pass through the reactor typically converts about 15% of the nitrogen and hydrogen to ammonia. However, unreacted gases are recycled, allowing for overall conversion rates of up to 97% in modern plants.

In conclusion, the optimal conditions for the Haber process - high temperature, high pressure, and the use of catalysts - are a testament to the practical application of chemical principles like Le Chatelier's principle in industrial settings. By carefully manipulating these conditions, engineers have turned a challenging reaction into one of the most important industrial processes of the modern world, producing the ammonia necessary for fertilizers that feed billions. Understanding these principles not only illuminates the Haber process but also provides a framework for analyzing and optimizing other chemical equilibria in both industrial and laboratory settings.

Industrial Application and Optimization of the Haber Process

The Haber process, also known as the Haber-Bosch process, is a cornerstone of modern industrial chemistry, playing a crucial role in the production of ammonia for fertilizers and other industrial applications. This process, developed by Fritz Haber and Carl Bosch in the early 20th century, has been continuously refined and optimized for industrial use, balancing efficiency, cost, and environmental considerations.

In industrial settings, the Haber process is applied on a massive scale. Large chemical plants utilize this process to produce millions of tons of ammonia annually. The basic principle involves combining nitrogen from the air with hydrogen, typically derived from natural gas, under high pressure and temperature conditions in the presence of an iron catalyst. While the chemistry remains fundamentally the same, the industrial application faces numerous challenges in optimization.

One of the primary challenges in optimizing the Haber process is energy consumption. The reaction requires temperatures around 400-450°C and pressures of 150-300 atmospheres, demanding significant energy input. Industries constantly strive to improve energy efficiency through various means. For instance, some plants employ advanced heat recovery systems to capture and reuse thermal energy from the process. Others have invested in more efficient catalysts that can operate at lower temperatures, reducing overall energy requirements.

Cost considerations are another critical factor in the industrial application of the Haber process. The high energy demands translate directly into operational costs, making energy efficiency not just an environmental concern but also an economic imperative. Additionally, the cost of raw materials, particularly natural gas for hydrogen production, significantly impacts the overall economics of ammonia production. To address this, some industries are exploring alternative hydrogen sources, such as electrolysis powered by renewable energy, though this approach is still in its early stages of industrial adoption.

The environmental impact of the Haber process is substantial, primarily due to its high energy consumption and the associated greenhouse gas emissions. The process is estimated to account for about 1-2% of global energy consumption and CO2 emissions. Industries are increasingly under pressure to reduce this environmental footprint. Some approaches include carbon capture and storage technologies, which aim to trap and sequester CO2 emissions from the process. Others are investigating the use of green hydrogen, produced through electrolysis using renewable energy, to make the process more sustainable.

Balancing these factors - energy efficiency, cost, and environmental impact - is a complex challenge for industries. Many companies employ sophisticated process control systems and modeling tools to optimize their operations continuously. For example, BASF, one of the world's largest chemical companies, has implemented advanced process control systems in its ammonia plants, resulting in significant improvements in energy efficiency and product yield.

Real-world examples illustrate the ongoing efforts to optimize the Haber process. In Norway, Yara International has been working on a project to produce "green ammonia" using hydrogen from electrolysis powered by renewable energy. This approach aims to significantly reduce the carbon footprint of ammonia production. In the United States, CF Industries has announced plans to implement carbon capture and storage at several of its ammonia production facilities, targeting a substantial reduction in CO2 emissions.

The optimization of the Haber process also extends to catalyst development. Researchers and industries are continually working on improving catalyst efficiency and longevity. For instance, Haldor Topsoe, a leading catalyst company, has developed advanced catalysts that allow for lower operating pressures in ammonia synthesis, contributing to energy savings and reduced equipment costs.

In conclusion, the industrial application of the Haber process represents a delicate balance between technological capability, economic viability, and environmental responsibility. As global demand for ammonia continues to grow, driven by agricultural needs and emerging applications like hydrogen storage, the pressure to optimize this process further intensifies. The future of the Haber process in industry will likely see continued innovations in energy efficiency, sustainable hydrogen production, and environmental mitigation strategies, as the sector strives to meet the world's ammonia needs in an increasingly sustainable manner.

Impact of the Haber Process on Agriculture and Society

The Haber process, developed in the early 20th century, has had profound implications for agriculture and global food production. This revolutionary method of synthesizing ammonia from nitrogen and hydrogen has dramatically increased the availability of nitrogen fertilizers, transforming modern farming practices and food security worldwide. The process, also known as the Haber-Bosch process, has been hailed as one of the most significant scientific achievements of the 20th century, with far-reaching consequences for human society.

Prior to the Haber process, agriculture relied heavily on natural sources of nitrogen, such as animal manure and crop rotation with legumes. These methods were often insufficient to meet the growing demand for food as the global population expanded. The introduction of synthetic nitrogen fertilizers derived from the Haber process changed this landscape dramatically. Farmers could now apply readily available, concentrated nitrogen to their crops, leading to substantial increases in crop yields.

Statistics illustrate the remarkable impact of the Haber process on food production. It is estimated that synthetic nitrogen fertilizers, made possible by this process, are responsible for sustaining approximately 40% of the world's population. Crop yields have increased significantly since the widespread adoption of these fertilizers. For instance, global cereal production has more than tripled since the 1960s, despite only a 30% increase in cultivated land area. This dramatic increase in productivity has been crucial in feeding the growing world population, which has more than doubled during the same period.

The positive impacts of the Haber process on food security cannot be overstated. The increased availability of nitrogen fertilizers has allowed farmers to produce more food on the same amount of land, reducing the pressure to convert natural habitats into farmland. This has been particularly important in developing countries, where improved crop yields have helped alleviate hunger and malnutrition. The Green Revolution of the mid-20th century, which significantly increased agricultural production in many parts of the world, was largely dependent on the availability of synthetic fertilizers made possible by the Haber process.

However, the widespread use of nitrogen fertilizers has also led to environmental concerns. Excessive application of these fertilizers can result in nutrient runoff, causing water pollution and eutrophication of water bodies. This can lead to algal blooms, fish kills, and degradation of aquatic ecosystems. Additionally, the production of synthetic fertilizers through the Haber process is energy-intensive, contributing to greenhouse gas emissions and climate change. It is estimated that the Haber process accounts for about 1-2% of global energy consumption.

The environmental impact of nitrogen fertilizers has prompted research into more sustainable agricultural practices. Precision agriculture techniques, which optimize fertilizer application based on specific crop needs and soil conditions, are being developed to reduce excess use. There is also growing interest in organic farming methods and the use of cover crops to naturally fix nitrogen in the soil.

Despite these challenges, the historical context of the Haber process reveals its immense significance in modern society. Before its invention, many experts predicted widespread famine due to the inability to produce enough food for the growing population. The process not only averted this crisis but also fundamentally changed the nature of agriculture and food production. It has allowed for the intensification of agriculture, freeing up labor for other sectors of the economy and contributing to urbanization and industrialization.

In conclusion, the Haber process has had a transformative impact on agriculture and global food production. Its role in increasing crop yields and enhancing food security has been crucial in supporting the world's growing population. While the environmental concerns associated with nitrogen fertilizers cannot be ignored, ongoing research and technological advancements aim to mitigate these issues. As we move forward, balancing the benefits of increased food production with environmental sustainability remains a key challenge for agriculture in the 21st century.

Future Developments and Alternatives

The Haber process, while revolutionary in its time, has been the subject of ongoing research and development aimed at improving efficiency and sustainability. Current research in this field is focused on addressing the environmental concerns and energy intensity associated with traditional ammonia production. Scientists and engineers are exploring various avenues for future developments, including alternative methods and more sustainable approaches to ammonia synthesis.

One promising area of research involves the use of renewable energy sources to power the Haber process. As the world shifts towards cleaner energy, integrating solar, wind, or hydroelectric power into ammonia production could significantly reduce its carbon footprint. Some pilot projects are already underway, demonstrating the feasibility of using green hydrogen (produced through electrolysis powered by renewable energy) as a feedstock for ammonia synthesis.

Another exciting development is the exploration of biological nitrogen fixation as an alternative to the Haber process. Researchers are studying the enzymes and bacteria that naturally fix nitrogen in soil, aiming to replicate or enhance these processes for industrial-scale ammonia production. This bio-inspired approach could potentially operate at ambient temperatures and pressures, dramatically reducing energy requirements.

Advancements in catalyst technology are also at the forefront of Haber process improvements. Scientists are investigating novel materials and nanostructures that could increase reaction rates and selectivity while reducing the need for extreme conditions. Some promising candidates include ruthenium-based catalysts and metal-organic frameworks (MOFs), which offer high surface areas and tunable properties.

The concept of decentralized, small-scale ammonia production is gaining traction as well. This approach could reduce transportation costs and emissions associated with large-scale centralized production. Modular and portable ammonia synthesis units are being developed, which could be particularly beneficial for remote agricultural areas or regions with limited infrastructure.

Electrochemical ammonia synthesis is another alternative method under investigation. This process aims to produce ammonia directly from nitrogen and water using electricity, potentially eliminating the need for hydrogen gas as a separate feedstock. While still in its early stages, this technology holds promise for more efficient and flexible ammonia production.

As research progresses, the principles of green chemistry are increasingly being applied to ammonia synthesis. This includes exploring room-temperature reactions, using abundant and non-toxic catalysts, and developing closed-loop systems that minimize waste and maximize resource efficiency.

Students and researchers in this field are encouraged to consider the following critical questions: What are the main challenges in scaling up these alternative methods? How can we balance the need for increased food production with environmental sustainability? What role might policy and economic incentives play in driving innovation in ammonia synthesis?

The future of ammonia production presents both challenges and opportunities. While the Haber process remains the dominant method, the push for more sustainable and environmentally friendly approaches is driving innovation. As we continue to explore alternative methods and improve existing processes, the goal is to develop ammonia production techniques that are not only efficient but also align with the principles of sustainable chemistry and contribute to a greener future for agriculture and industry.

Conclusion

The Haber process stands as a cornerstone of modern chemistry and agriculture, revolutionizing ammonia synthesis and fertilizer production. This industrial chemistry breakthrough has had far-reaching impacts on global food security and industrial applications. The introduction video provided a crucial foundation for understanding the process's intricacies. As we've explored, the Haber process exemplifies the intersection of scientific innovation and practical application, demonstrating how chemical reactions can be harnessed to address global challenges. Its significance in feeding the world's growing population cannot be overstated. We encourage readers to delve deeper into this fascinating subject, perhaps by exploring related topics such as catalysis or nitrogen fixation. For those intrigued by the Haber process, consider investigating its environmental implications or the ongoing research to improve its efficiency. We invite you to share your thoughts or questions about the Haber process in the comments section below, fostering further discussion on this pivotal chemical reaction.

Equilibrium in Real Life Chemistry: The Haber Process Introduction

Equilibrium in real life chemistry. The Haber process: introduction.

Step 1: Introduction to the Haber Process

The Haber process, also known as the Haber-Bosch process, is a critical industrial method for synthesizing ammonia from nitrogen and hydrogen gases. This process is a practical application of chemical equilibrium principles in real-life industry. Developed by Fritz Haber and Karl Bosch in the early 1900s, the process has become essential for producing ammonia, which is a key component in fertilizers. The increased crop yields from these fertilizers support a significant portion of the world's population, estimated to be between 15% to 20%.

Step 2: Understanding the Chemical Reaction

The chemical equation for the Haber process is as follows:

N2(g) + 3H2(g) 2NH3(g)

This reaction involves nitrogen gas (N2) and hydrogen gas (H2) reacting to form ammonia (NH3). The reaction is exothermic, meaning it releases heat, and an equilibrium is established between the reactants and the product.

Step 3: Historical Context and Importance

Initially, the Haber process was developed to produce nitrates for explosives, particularly before World War I. However, its primary use today is for producing ammonia for fertilizers. The process has been in use since the 1940s and has grown in scale, with nearly 200 million tonnes of ammonia produced annually. This production is crucial for meeting the food demands of a growing global population.

Step 4: The Role of Equilibrium in the Haber Process

The Haber process is a prime example of how equilibrium concepts are applied in industrial chemistry. The reaction conditions, such as temperature and pressure, significantly affect the equilibrium position and the yield of ammonia. According to Le Chatelier's principle, if a system at equilibrium is disturbed, it will adjust to counteract the disturbance and restore equilibrium.

Step 5: Impact of Temperature on the Reaction

Since the Haber process is exothermic, increasing the temperature will shift the equilibrium towards the reactants, reducing the yield of ammonia. This is because the system will try to absorb the added heat by favoring the endothermic reverse reaction. Therefore, a lower temperature is preferred to maximize ammonia production, but it must be balanced with the need for a reasonable reaction rate.

Step 6: Optimizing Reaction Conditions

To optimize the Haber process, a compromise between temperature, pressure, and the use of a catalyst is necessary. High pressure favors the formation of ammonia because it shifts the equilibrium towards the side with fewer gas molecules (ammonia). However, extremely high pressures can be costly and pose safety risks. A catalyst, typically iron with potassium hydroxide, is used to increase the reaction rate without affecting the equilibrium position.

Step 7: Practical Applications and Global Impact

The ammonia produced through the Haber process is vital for manufacturing fertilizers, which in turn support global agriculture. The increased crop yields from these fertilizers help feed a significant portion of the world's population. The process's efficiency and scalability make it a cornerstone of modern industrial chemistry and food production.

Conclusion

The Haber process exemplifies the application of equilibrium principles in real-life chemistry. By understanding and manipulating reaction conditions, chemists can optimize the production of ammonia, a crucial compound for global agriculture. The process's historical development, chemical principles, and practical applications highlight its importance in both industry and society.

FAQs

Here are some frequently asked questions about the Haber process:

  1. What is the equation for the Haber process?

    The equation for the Haber process is: N + 3H 2NH. This represents the reversible reaction between nitrogen and hydrogen to produce ammonia.

  2. What are the optimal conditions for the Haber process?

    The optimal conditions for the Haber process typically involve temperatures of 400-450°C, pressures of 150-300 atmospheres, and the use of an iron-based catalyst. These conditions balance reaction rate, yield, and practical considerations.

  3. Why is the Haber process important?

    The Haber process is crucial for producing ammonia, which is essential for fertilizer production. It has dramatically increased agricultural productivity, supporting global food security and population growth.

  4. Is the Haber process exothermic or endothermic?

    The Haber process is exothermic, meaning it releases heat. This characteristic influences the optimal conditions for the reaction, as higher temperatures favor the reverse reaction according to Le Chatelier's principle.

  5. What are the environmental concerns associated with the Haber process?

    The main environmental concerns include high energy consumption, contributing to greenhouse gas emissions, and the potential for nitrogen runoff from fertilizers, which can lead to water pollution and ecosystem disruption.

Prerequisite Topics for Understanding the Haber Process

When delving into the complexities of the Haber process, a fundamental understanding of certain prerequisite topics is crucial. Two key areas that provide a solid foundation for comprehending this important industrial process are activation energy and introduction to kinetics.

The Haber process, also known as the Haber-Bosch process, is a method for synthesizing ammonia from nitrogen and hydrogen gases. This process is of immense industrial importance, as it is the primary method of producing fertilizers that support global food production. To truly grasp the intricacies of this process, one must first understand the concept of activation energy.

Activation energy refers to the minimum energy required for a chemical reaction to occur. In the context of the Haber process, understanding the activation energy barrier is crucial. The reaction between nitrogen and hydrogen to form ammonia requires a significant amount of energy to break the strong triple bond in nitrogen molecules. By grasping the concept of activation energy, students can better comprehend why high temperatures and pressures are necessary for the Haber process to proceed efficiently.

Equally important is an introduction to kinetics. Reaction kinetics deals with the rates of chemical reactions and the factors that influence these rates. In the Haber process, understanding kinetics is essential for optimizing the yield of ammonia. Factors such as temperature, pressure, and the presence of catalysts all play crucial roles in determining the rate and efficiency of the reaction.

By studying reaction kinetics, students can gain insights into why the Haber process operates under specific conditions. For instance, the use of an iron catalyst in the process can be better understood through the lens of kinetics, as it lowers the activation energy and increases the rate of reaction without being consumed.

Moreover, the principles of kinetics help explain the reversible nature of the Haber process and the concept of chemical equilibrium. This knowledge is crucial for understanding how adjusting factors like pressure can shift the equilibrium to favor the production of ammonia.

In conclusion, a solid grasp of activation energy and kinetics provides the necessary foundation for understanding the Haber process. These prerequisite topics offer valuable insights into the energy requirements, reaction rates, and optimization strategies employed in this vital industrial process. By mastering these fundamental concepts, students can develop a more comprehensive and nuanced understanding of the Haber process and its significance in modern chemistry and industry.

In this lesson, we will learn:

  • To recall the Haber process and the chemicals involved.
  • How to explain how changing reaction conditions affects the Haber process.
  • How to explain the optimum conditions of the Haber process using Le Chatelier's principle.

Notes:

  • Some very important chemical processes happen in an equilibrium. One example is the Haber process, which converts nitrogen and hydrogen gas into ammonia, a common fertilizer now used to increase crop yield.

    N2(g)+3H2(g) 2NH3(g)\mathrm{N_{2(g)} + 3H_{2(g)} \ \rightleftharpoons 2NH_{3(g)}}

    The Haber process was invented by a German chemist called Fritz Haber in the early 1900's as a way to get to nitrates which were needed to make explosives at the time. Today, over a hundred million tons of ammonia are made every year, and the figure has steadily increased since the 1940s.
    The Haber process is an exothermic reaction that happens in equilibrium, which creates the following problem: increasing the temperature will make the reaction faster, but will shift the equilibrium to make less products!
  • As we saw in C12.1.9: Entropy and spontaneous reactions, when enthalpy and entropy favor opposite sides of a reaction, an equilibrium is usually the result. Chemical companies want and need to run the Haber process to make as much ammonia as possible using the least resources as possible. This means:
    • Make it cheap to run, using less heat, pressure, energy and chemicals.
    • Make it low-maintenance, so things don't need replacing, processing or support to work properly.
    • Do it quickly, with the fastest reaction rate possible.
  • Using our knowledge of equilibrium, what can we do to the reaction conditions to make as much ammonia as possible?
    • Increasing the operating temperature…
      • …will increase the rate of reaction, which will speed up both the forward and the reverse reaction.
      • …will shift the equilibrium to the left, which favors the reverse reaction and makes more reactants. The reverse reaction is endothermic and will absorb the added heat from the increased temperature.
    • Decreasing the operating temperature…
      • …will lower the rate of reaction, slowing both forward and reverse reactions and the time taken for the process to reach equilibrium.
      • …will shift the equilibrium to the right, making more products by favoring the forward reaction. Favoring the exothermic forward reaction releases heat, countering the decreasing temperature.
    • Increasing the pressure…
      • …will shift the equilibrium to the right to favor the products. There are less moles of gas in the products, so this counters the increased pressure applied to the system.
      • …will be extremely expensive and quite dangerous for a reaction running on a large scale!
    • Decreasing the pressure…
      • Will shift the equilibrium to the left to favor the reactants to counter the decrease in pressure.
    What else could be added to the equilibrium mixture to make the forward reaction (or any reaction!) run faster?
  • Resolving these issues, chemical industry uses compromise conditions to extract the most ammonia from this process:
    • Temperature is set at around 450°C, which is high enough for a high rate of reaction without excessively favoring the reactants in the equilibrium.
    • Pressure is around 200 atmospheres (200 times higher than normal air pressure). This is high enough to favor the products in the reaction, but not too high that it is extremely expensive or dangerous to run machinery.
    • An iron catalyst is added. This speeds up the rate of the reaction both ways and decreases the time taken to get to equilibrium.