Skip to main content
Biology LibreTexts

Lecture 08: Redox reactions

  • Page ID
    16897
  •  

    \( \newcommand{\vecs}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \) \( \newcommand{\vecd}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash {#1}}} \)\(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\) \(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\)\(\newcommand{\AA}{\unicode[.8,0]{x212B}}\)

    Metabolism in BIS2A

    Cellular metabolism represents roughly 1/3 of the BIS2A curriculum. While this may seem like a lot, we cover very little of what a classic course in metabolism (like BIS103) will cover, and an even smaller (really minuscule) fraction of the metabolism that occurs on the planet. What we do cover, however, is very important foundational knowledge. You will learn about some common chemical transformations that are associated with the transformation of life's molecular building blocks and about different core modes of energy transfer that you will encounter often in biology. The energy story and the design challenge rubrics introduced earlier will become increasingly important in these next few modules and beyond.

    What have we learned? How will it relate to metabolism?

    1. We have focused on the identification and chemical properties of common biological functional groups. As we dive into metabolism, this will help you be familiar with and sometimes even predict the chemical nature/reactivity of compounds you have never seen before.
    2. We have practiced recognizing and classifying molecules into four major functional groups. This will help you as we begin to discuss how to build and break down these molecules.
    3. We have learned some basic thermodynamics, giving us a common set of concepts with which to discuss whether a biochemical reaction or process is likely to occur, and if so in which direction and how fast. This will be critical as we begin to consider some of the key types of reactions that take place in metabolism.
    4. We have learned and practiced the energy story rubric. This too will allow us to systematically examine new biochemical reactions and processes and to discuss them with a common language and approach that is consistent and reinforces the lessons we learned about thermodynamics.

    A brief overview of this section

    • You will be introduced to an important concept called reduction potential and you will be given the opportunity to use a redox tower. There is also a discussion on redox chemistry in your discussion manual. Make sure to use both resources.
    • You will be introduced to two major players in metabolism, ATP and NADH. You will be expected to recognize their structures if shown on an exam.
    • The metabolic pathway glycolysis will be covered in detail. Keep in mind that we want you to be able to look at any reaction and tell us an energy story of that reaction. By no means should you spend time trying to memorize these pathways (though it will help tremendously to remember some big picture things - these will be stressed). Often we will give you the pathway as a figure on the exams. Glycolysis ultimately produces 2 ATP via a process called substrate level phosphorylation, 2 NADH and 2 pyruvate compounds.
    • We will use the reactions of the TCA cycle to create multiple examples of energy stories. The TCA cycle will also produce more ATP, NADH and completely oxidize glucose into CO2.
    • We will look at an alternative pathway to that of the TCA cycle, fermentation. In fermentation for the first time we will see NADH used as a reactant in a metabolic reaction.
    • We will follow NADH to the end of its journey, as it donates its electrons to the electron transport chain (ETC). In this module you will need to be able to use a redox tower. The ETC produces a proton gradient. No ATP is directly generated in this process. However, the proton gradient is then used by the cell (among other things) to run an enzyme called ATP synthase which catalyzes the reaction ADP + Pi --> ATP. This method of ATP production (called oxidative respiration) results in much more ATP being produced than substrate level phosphorylation.
    • And finally, we will go through the process of photosynthesis.

    Reduction-Oxidation Reactions

    In this class we are going to focus on redox reactions that are biologically associated. The majority of the reactions we discuss occur in the context of metabolic pathways (connected sets of metabolic reactions) where compounds may be consumed by the cell, broken down into smaller parts and then reassembled into larger macromolecules.

    Lets start with some generic reactions

    Transferring electrons between two compounds results in one of these compounds loosing an electron, and one of the compounds gaining an electron. For example, look at the figure below. If we use the energy story rubric to look at the overall reaction we can compare the before and after characteristics of the reactants and products. What happens to the matter (stuff) before and after the reaction? Compound A starts as neutral and becomes positively charged. Compound B starts as neutral and becomes negatively charged. Because electrons are negatively charged, we can follow the movement of electrons from compound A to B by looking at the change in charge. A looses an electron (becoming positively charged), and in so doing we say that A has become oxidized. Oxidation is associated with the loss of electron(s). B gains the electron (becoming negatively charged), and we say that B has become reduced. Reduction is associated with gain of electrons. We also know, since something happened that energy must have been either transferred and/or reorganized in this process and we'll consider this shortly.

    In Bis2A we expect you to become familiar with this terminology. Try to learn it and learn to use it as soon as possible - we will use the terms frequently and will not have the time to define them each time.

    generic.jpg

    Figure 1. A generic redox reaction. The full reaction is A +B goes to A+ + B-. The two half reactions are shown in the blue box. A is oxidized by the reaction and B is reduced by the reaction.

    When an electron(s) is lost, or a molecule is oxidized, the electron(s) must then passed to another molecule. The molecule gaining the electron is said to be reduced. The oxidation and reduction reactions are always paired in what is known as an oxidation-reduction reaction (also called a red/ox reaction).

    For emphasis: In Bis2a we expect you to become familiar with this terminology. Try to learn it and learn to use it as soon as possible - we will use the terms frequently and will not have the time to define them each time.

    Remember the Definitions:

    Screen Shot 2017-01-08 at 7.04.42 PM.png

    The Half Reaction

    To formalize our common understanding of red/ox reactions, we introduce the concept of the half reaction. Two half reactions are required to make the full red/ox reaction. Each half reaction can be thought of as a description of what happens to one of the two molecules involved in the red/ox reaction. This is illustrated below. In this example compound AH is being oxidized by compound B+; electrons are moving from AH to B+ to generate A+ and BH. Each reaction can be thought of as two half reactions: Where AH is being oxidized and a second reaction where B+ is being reduced to BH. These two reactions are considered coupled, a term that indicates that these two reactions occur together, at the same time.

    figure1.jpg

    Figure 2. Generic red/ox reaction where compound AH is being oxidized by compound B+. Each half reaction represents a single species or compound to either lose or gain electrons (and a subsequent proton as shown in the figure above). In half reaction #1 AH loses a proton and 2 electrons: in the second half reaction, B+ gains 2 electrons and a proton. In this example HA is oxidized to A+ while B+ is reduced to BH.

    Possible discussion

    If you consider a generic redox reaction and reflect back on the thermodynamic lectures what factor will determine whether a redox reaction will "go" in a particular direction spontaneously and what might determine its rate?

    Reduction Potential

    By convention we analyze and describe red/ox reactions with respect to reduction potentials, a term that quantitatively describes the "ability" of a compound to gain electrons. This value of the reduction potential is determined experimentally but for the purpose of this course we assume that the reader will accept that the reported values are reasonably correct. We can anthropomorphize the reduction potential by saying that it is related to the strength with which a compound can “attract” or “pull” or “capture” electrons. Not surprisingly this is is related to but not identical to electronegativity.

    What is this intrinsic property to attract electrons?

    Different compounds, based on their structure and atomic composition have intrinsic and distinct attractions for electrons. This quality is termed reduction potential or E0’and is a relative quantity (relative by comparison to some “standard” reaction). If a test compound has a stronger "attraction" to electrons than the standard (if the two competed the test compound would "take" electrons from the standard compound), we say that the test compound has a positive reduction potential whose magnitude is proportional to how much more it "wants" electrons than the standard compound. The relative strength of the compound in comparison to the standard is measured and reported in units of Volts (V)(sometimes written as electron volts or eV) or milliVolts (mV). The reference compound in most redox towers is H2.

    Possible discussion

    Rephrase for yourself: How do you describe or think about the difference between the concept of electronegativity and redox potential?

    The Redox Tower

    All kinds of compounds can participate in redox reactions. A tool has been developed to graphically tabulate redox half reactions based on their E0' values and to help us predict the direction of electron flow between potential electron donors and acceptors. Whether a particular compound can act as an electron donor (reductant) or electron acceptor (oxidant) depends critically on what other compound it is interacting with. The electron tower usually ranks a variety of common compounds (their half reactions) from most negative E0', compounds that readily get rid of electrons, to the most positive E0', compounds most likely to accept electrons. In addition, each half reaction is written by convention with the oxidized form on the left/followed by the reduced form on the right of the slash.
    For example the half reaction for the reduction of NAD+ to NADH is written:
    NAD+/NADH. In the tower below, the number of electrons that are transferred is also listed. For example the reduction of NAD+ to NADH involves two electrons, written in the table as 2e-.

    An electron tower is shown below.

    figure9.jpg

    Figure 3. Common Red/ox tower used in Bis2A. By convention the tower half reactions are written with the oxidized form of the compound on the left and the reduced form on the right. Compounds that make excellent electron donors are found at the top of the tower. Compounds such as Glucose and Hydrogen gas are excellent electron donors. Notice, that they are found on the right hand side of the red/ox pair half reactions. At the other end of the tower lies compounds that make excellent terminal electron acceptors, such as Oxygen and Nitrite, these compounds are found on the left side of the red/ox pair and have a positive E0' value.

    Video on electron tower

    For a short video on how to use the electron tower in red/ox problems click here. This video was made by Dr. Easlon for Bis2A students. (This is quite informative.)

    What is the relationship between ΔE0' and ΔG?

    The question now becomes: how do we know if any given redox reaction is energetically spontaneous or not (exergonic or endergonic) and regardless of direction, what the free energy difference is? The answer lies in the difference in the reduction potentials of the two compounds. The difference in the reduction potential for the reaction or E0' for the reaction, is the difference between the E0' for the oxidant (the compound getting the electrons and causing the oxidation of the other compound) and the reductant (the compound losing the electrons). In our generic example below, AH is the reductant and B+ is the oxidant. Electrons are moving from AH to B+. Using the E0' of -0.32 for the reductant and 0.82 for the oxidant the total change in E0' or ΔE0' is 1.14 eV.

    figure11.jpg

    Figure 4. Generic redox reaction with half reactions written with reduction potential (E0') of the two half reactions indicated.

    The change in ΔE0' correlates to changes in Gibbs free energy, ΔG. In general a large positive ΔE0' is proportional to a large negative ΔG. The reactions are exergonic and spontaneous. For a reaction to be exergonic the reaction needs to have a negative change in free energy or -ΔG, this will correspond to a positive ΔE0'. In other words, when electrons flow "downhill" in a redox reaction from a compound with a higher (more positive) reduction potential to a second compound with a lower (less positive) reduction potential, they release free energy. The greater the voltage, ΔE0', between the two components, the greater the energy available when electron flow occurs. It is, in fact, possible to quantify the amount of free energy available. The relationship is given by the Nernst equation:

    nernst.png

    Figure 5. The Nernst equation relates free energy of a redox reaction to the difference in reduction potential between the reduced products of the reaction and oxidized reactant.
    Attribution: Marc T. Facciotti

    Where:

    • n is the number of moles of electrons transferred
    • F is the Faraday constant of 96.485 kJ/V. Sometimes it is given in units of kcal/V which is 23.062 kcal/V, which is the amount of energy (in kJ or kcal) released when one mole of electrons passes through a potential drop of 1 volt

    Note

    What you should notice is that ΔG and ΔE have an inverse relationship: When ΔG is positive, ΔE is negative and when ΔG is negative, ΔE is positive. For additional review see the redox discussion in the Bis2A Discussion Manual.

    Glycolysis: an overview

    Organisms, whether unicellular or multicellular, need to find ways of getting at least two key things from their environment: (1) matter or raw materials for maintaining a cell and building new cells and (2) energy to help with the work of staying alive and reproducing. Energy and the raw materials may come from different places. For instance, organisms that primarily harvest energy from sunlight will get raw materials for building biomolecules from sources like CO2. By contract, some organisms rely on redox reactions with small molecules and/or reduced metals for energy and get their raw materials for building biomolecules from compounds unconnected to the energy source. Meanwhile, some organisms (including ourselves), have evolved to get energy AND the raw materials for building and cellular maintenance from sometimes associated sources.

    Glycolysis is the first metabolic pathway discussed in BIS2A; a metabolic pathway is a series of linked biochemical reactions. Because of its ubiquity in biology, it is hypothesized that glycolysis was probably one of the earliest metabolic pathways to evolve (more on this later). Glycolysis is a ten-step metabolic pathway that is centered on the processing of glucose for both energy extraction from chemical fuel and for the processing of the carbons in glucose into various other biomolecules (some of which are key precursors of many much more complicated biomolecules). Our study of glycolysis will therefore be examined using the precepts outlined in the energy challenge rubric that ask us to formally consider what happens to BOTH matter and energy in this multistep process.

    The energy story and design challenge

    Our investigation of glycolysis is a good opportunity to examine a biological process using both the energy story and the design challenge rubrics and perspectives.

    The design challenge rubric will try to get you to think actively, and broadly and specifically, about why we are studying this pathway—what is so important about it? What "problems" does the evolution of a glycolytic pathway allow life to solve or overcome? We will also want to think about alternate ways to solve the same problems and why they may or may not have evolved. Later, we will examine a hypothesis for how this pathway—and other linked pathways—may have actually evolved, and thinking about alternative strategies for satisfying various constraints will come in handy then.

    In the context of the energy story, we will ask you to think about glycolysis as a process from which something can be learned by analyzing what happens to both matter and energy. That is, even though it is a ten-step biochemical pathway, we propose that some insight can be learned by carefully examining the process as a set of matter and energy inputs and outputs, a process with a beginning and an end.

    So what is glycolysis? Let's start to find out.

    glycolysis_2panel.png

    Figure 1. The ten biochemical reactions of glycolysis are shown. Enzymes are labeled in blue. The structure of each sugar-derived compound is depicted as a molecular model; other reactants and products may be abbreviated (e.g., ATP, NAD+, etc.). The box surrounding the reaction catalyzed by glyceraldehyde 3-phosphate dehydrogenase indicates that this reaction is of special interest in the course. Attribution: Marc T. Facciotti (original work)

    Table 1. This table shows glycolytic enzymes and measurements of the energy at standard state (ΔG°'/(kJ/mol)) compared with measurements taken from a living cell (ΔG/(kJ/mol)). Under conditions of constant temperature and pressure, (ΔG°'/(kJ/mol)), reactions will occur in the direction that leads to a decrease in the value of the Gibbs free energy. Cellular measurements of ΔG can be dramatically different than ΔG°' measurements due to cellular conditions, such as concentrations of relevant metabolites etc. There are three large, negative ΔG drops in the cell in the process of glycolysis. These reactions are considered irreversible and are often subject to regulation.

    Enzyme Step ΔG/(kJ/mol) ΔG°'/(kJ/mol)
    Hexokinase 1 -34 -16.7
    Phosphoglucose isomerase 2 -2.9 1.67
    Phosphofructokinase 3 -19 -14.2
    Fructose-bisphosphate aldolase 4 -0.23 23.9
    Triose phosphate isomerase 5 2.4 7.56
    Glyceraldehyde 3-phosphate dehydrogenase 6 -1.29 6.30
    Phosphoglycerate kinase 7 0.09 -18.9
    Phosphoglycerate mutase 8 0.83 4.4
    Enolase 9 1.1 1.8
    Pyruvate kinase 10 -23.0 -31.7

    Overall, the glycolytic pathway consists of 10 enzyme-catalyzed steps. The primary input into this pathway is a single molecule of glucose, though we will discover that molecules may feed in and out of this pathway at various steps. We will focus our attention on (1) consequences of the overall process, (2) several key reactions that highlight important types of biochemistry and biochemical principles we will want to carry forward to other contexts, and (3) alternative fates of the intermediates and products of this pathway.

    Note for reference that glycolysis is an anaerobic process; there is no requirement for molecular oxygen in glycolysis (oxygen gas is not a reactant in any of the chemical reactions in glycolysis). Glycolysis occurs in the cytosol or cytoplasm of cells. For a short (three-minute) overview YouTube video of glycolysis, click here.

    First half of glycolysis: energy investment phase

    The first few steps of glycolysis are typically referred to as an "energy investment phase" of the pathway. This, however, doesn't make much intuitive sense (in the framework of a design challenge; it's not clear what problem this energy investment solves) if one only looks at glycolysis as an "energy-producing" pathway and until these steps of glycolysis are put into a broader metabolic context. We'll try to build that story as we go, so for now just recall that we mentioned that some of the first steps are often associated with energy investment and ideas like "trapping" and "commitment" that are noted in the figure below.

    Step 1 of glycolysis:

    The first step in glycolysis, shown below in Figure 2, is glucose being catalyzed by hexokinase, an enzyme with broad specificity that catalyzes the phosphorylation of six-carbon sugars. Hexokinase catalyzes the phosphorylation of glucose, where glucose and ATP are substrates for the reaction, producing a molecule called glucose 6-phosphate and ADP as products.

    glycolysis_first_half.png

    Figure 2. The first half of glycolysis is called the energy investment phase. In this phase, the cell expends two ATPs into the reactions. Attribution: Marc T. Facciotti (original work)

    Suggested discussion

    The paragraph above states that the enzyme hexokinase has "broad specificity." This means that it can catalyze reactions with different sugars, not just glucose. From a molecular perspective, can you explain why this might be the case? Does this challenge your conception of enzyme specificity? If you Google the term "enzyme promiscuity" (don't worry; it's safe for work), does this give you a broader appreciation for enzyme selectivity and activity?

    The conversion of glucose to the negatively charged glucose 6-phosphate significantly reduces the likelihood that the phosphorylated glucose leaves the cell by diffusion across the hydrophobic interior of the plasma membrane. It also "marks" the glucose in a way that effectively tags it for several different possible fates (see Figure 3).

    fates_of_g6p.png

    Figure 3. Note that this figure indicates that glucose 6-phosphate can, depending on cellular conditions, be directed to multiple fates. While it is a component of the glycolytic pathway, it is not only involved in glycolysis but also in the storage of energy as glycogen (colored in cyan) and in the building of various other molecules like nucleotides (colored in red). Source: Marc T. Facciotti (original work)

    As Figure 3 indicates, glycolysis is but one possible fate for glucose 6-phosphate (G6P). Depending on cellular conditions, G6P may be diverted to the biosynthesis of glycogen (a form of energy storage), or it may be diverted into the pentose phosphate pathway for the biosynthesis of various biomolecules, including nucleotides. This means that G6P, while involved in the glycolytic pathway, is not solely tagged for oxidation at this phase. Perhaps showing the broader context that this molecule is involved in (in addition to the rationale that tagging glucose with a phosphate decreases the likelihood that it will leave the cell) helps to explain the seemingly contradictory (if you only consider glycolysis as an "energy-producing" process) reason for transferring energy from ATP onto glucose if it is only to be oxidized later—that is, glucose is not only used by the cell for harvesting energy and several other metabolic pathways depend on the transfer of the phosphate group.

    Step 2 of glycolysis:

    In the second step of glycolysis, an isomerase catalyzes the conversion of glucose 6-phosphate into one of its isomers, fructose 6-phosphate. An isomerase is an enzyme that catalyzes the conversion of a molecule into one of its isomers.

    Step 3 of glycolysis:

    The third step of glycolysis is the phosphorylation of fructose 6-phosphate, catalyzed by the enzyme phosphofructokinase. A second ATP molecule donates a phosphate to fructose 6-phosphate, producing fructose 1,6-bisphosphate and ADP as products. In this pathway, phosphofructokinase is a rate-limiting enzyme, and its activity is tightly regulated. It is allosterically activated by AMP when the concentration of AMP is high and when it is moderately allosterically inhibited by ATP at the same site. Citrate, a compound we'll discuss soon, also acts as a negative allosteric regulator of this enzyme. In this way, phosphofructokinase monitors or senses molecular indicators of the energy status of the cells and can in response act as a switch that turns on or off the flow of the substrate through the rest of the metabolic pathway depending on whether there is “sufficient” ATP in the system. The conversion of fructose 6-phosphate into fructose 1,6-bisphosphate is sometimes referred to as a commitment step by the cell to the oxidation of the molecule in the rest of the glycolytic pathway by creating a substrate for and helping to energetically drive the next highly endergonic (under standard conditions) step of the pathway.

    Suggested discussion

    We discussed allosteric regulation of an enzyme in earlier modules but did so in a context where the enzyme was "alone." Now let's consider the enzyme in the context of an extended metabolic pathway(s). Can you now express why allosteric regulation is functionally important and how it can be used to regulate the flow of compounds through a pathway? Try to express yourself.

    Step 4 of glycolysis:

    In the fourth step in glycolysis, an enzyme, fructose-bisphosphate aldolase, cleaves 1,6-bisphosphate into two three-carbon isomers: dihydroxyacetone phosphate and glyceraldehyde 3-phosphate.

    Second half: energy payoff phase

    If viewed in the absence of other metabolic pathways, glycolysis has thus far cost the cell two ATP molecules and produced two small, three-carbon sugar molecules: dihydroxyacetone phosphate (DAP) and glyceraldehyde 3-phosphate (G3P). When viewed in a broader context, this investment of energy to produce a variety of molecules that can be used in a variety of other pathways doesn't seem like such a bad investment.

    Both DAP and G3P can proceed through the second half of glycolysis. We now examine these reactions.

    18323310-q-07-650-a542d8629a-1478168682.png

    Figure 4. The second half of glycolysis is called the energy payoff phase. In this phase, the cell gains two ATP and two NADH compounds. At the end of this phase, glucose has become partially oxidized to form pyruvate. Attribution: Marc T. Facciotti (original work).

    Step 5 of glycolysis:

    In the fifth step of glycolysis, an isomerase transforms the dihydroxyacetone phosphate into its isomer, glyceraldehyde 3-phosphate. The six-carbon glucose has therefore now been converted into two phosphorylated three-carbon molecules of G3P.

    Step 6 of glycolysis:

    The sixth step is key and one from which we can now leverage our understanding of the several types of chemical reactions that we've studied so far. If you're energy focused, this is finally a step of glycolysis where some of the reduced sugar is oxidized. The reaction is catalyzed by the enzyme glyceraldehyde 3-phosphate dehydrogenase. This enzyme catalyzes a multistep reaction between three substrates—glyceraldehyde 3-phosphate, the cofactor NAD+, and inorganic phosphate (Pi)—and produces three products: 1,3-bisphosphoglycerate, NADH, and H+. One can think of this reaction as two reactions: (1) an oxidation/reduction reaction and (2) a condensation reaction in which an inorganic phosphate is transferred onto a molecule. In this particular case, the redox reaction, a transfer of electrons off of G3P and onto NAD+, is exergonic, and the phosphate transfer happens to be endergonic. The net standard free energy change hovers around zero—more on this later. The enzyme here acts as a molecular coupling agent to couple the energetics of the exergonic reaction to that of the endergonic reaction, thus driving both forward. This processes happens through a multistep mechanism in the enzyme's active site and involves the chemical activity of a variety of functional groups.

    It is important to note that this reaction depends upon the availability of the oxidized form of the electron carrier, NAD+. If we consider that there is a limiting pool of NAD+, we can then conclude that the reduced form of the carrier (NADH) must be continuously oxidized back into NAD+ in order to keep this step going. If NAD+ is not available, the second half of glycolysis slows down or stops.

    Step 7 of glycolysis:

    In the seventh step of glycolysis, catalyzed by phosphoglycerate kinase (an enzyme named for the reverse reaction), 1,3-bisphosphoglycerate transfers a phosphate to ADP, forming one molecule of ATP and a molecule of 3-phosphoglycerate. This reaction is exergonic and is also an example of substrate-level phosphorylation.

    Possible discussion

    If a transfer of a phosphate from 1,3-BPG to ADP is exergonic, what does that say about the free energy of hydrolysis of the phosphate from 1,3-BPG as compared to the free energy of hydrolysis of the terminal phosphate on ATP?

    Step 8 of glycolysis:

    In the eighth step, the remaining phosphate group in 3-phosphoglycerate moves from the third carbon to the second carbon, producing 2-phosphoglycerate (an isomer of 3-phosphoglycerate). The enzyme catalyzing this step is a mutase (isomerase).

    Step 9 of glycolysis:

    Enolase catalyzes the ninth step. This enzyme causes 2-phosphoglycerate to lose water from its structure; this is a dehydration reaction, resulting in the formation of a double bond that increases the potential energy in the remaining phosphate bond and produces phosphoenolpyruvate (PEP).

    Step 10 of glycolysis:

    The last step in glycolysis is catalyzed by the enzyme pyruvate kinase (the enzyme in this case is named for the reverse reaction of pyruvate’s conversion into PEP) and results in the production of a second ATP molecule by substrate-level phosphorylation and the compound pyruvic acid (or its salt form, pyruvate). Many enzymes in enzymatic pathways are named for the reverse reactions, since the enzyme can catalyze both forward and reverse reactions (these may have been described initially by the reverse reaction that takes place in vitro, under non-physiological conditions).

    Outcomes of glycolysis

    Here are a couple of things to consider:

    One of the clear outcomes of glycolysis is the biosynthesis of compounds that can enter into a variety of metabolic pathways. Likewise, compounds coming from other metabolic pathways can feed into glycolysis at various points. So, this pathway can be part of a central exchange for carbon flux within the cell.

    If glycolysis is run long enough, the constant oxidation of glucose with NAD+ can leave the cell with a problem: how to regenerate NAD+ from the two molecules of NADH produced. If the NAD+ is not regenerated, all of the cell's NAD will be nearly completely transformed into NADH. So how do cells regenerate NAD+?

    Pyruvate is not completely oxidized; there is still some energy to be extracted. How might this happen? Also, what should the cell do with all of that NADH? Is there any energy there to extract?

    Strongly suggested discussion/exercise

    Can you write an energy story for the overall process of glycolysis? For energy terms, just worry about describing things in terms of whether they are exergonic or endergonic. When I say "overall process," I mean overall process: glucose should be listed on the reactant side of the arrow, and pyruvate should be listed on the product side of the arrow.

    Substrate-level phosphorylation (SLP)

    The simplest route to synthesize ATP is substrate-level phosphorylation. ATP molecules are generated (that is, regenerated from ADP) as a direct result of a chemical reaction that occurs in catabolic pathways. A phosphate group is removed from an intermediate reactant in the pathway, and the free energy of the reaction is used to add the third phosphate to an available ADP molecule, producing ATP. This very direct method of phosphorylation is called substrate-level phosphorylation. It can be found in a variety of catabolic reactions, most notably in two specific reactions in glycolysis (which we will discuss specifically later). Suffice it to say that what is required is a high-energy intermediate whose oxidation is sufficient to drive the synthesis of ATP.

    sub.jpg

    Figure 5. Here is one example of substrate-level phosphorylation occurring in glycolysis. There is a direct transfer of a phosphate group from the carbon compound onto ADP to form ATP. Attribution: Marc T. Facciotti (own work)

    In this reaction, the reactants are a phosphorylated carbon compound called G3P (from step 6 of glycolysis) and an ADP molecule, and the products are 1,3-BPG and ATP. The transfer of the phosphate from G3P to ADP to form ATP in the active site of the enzyme is substrate-level phosphorylation. This occurs twice in glycolysis and once in the TCA cycle (for a subsequent reading).

    Fermentation and regeneration of NAD+

    Section summary

    This section discusses the process of fermentation. Due to the heavy emphasis in this course on central carbon metabolism, the discussion of fermentation understandably focuses on the fermentation of pyruvate. Nevertheless, some of the core principles that we cover in this section apply equally well to the fermentation of many other small molecules.

    The purpose of fermentation

    The oxidation of a variety of small organic compounds is a process that is utilized by many organisms to garner energy for cellular maintenance and growth. The oxidation of glucose via glycolysis is one such pathway. Several key steps in the oxidation of glucose to pyruvate involve the reduction of the electron/energy shuttle NAD+ to NADH. At the end of section 5.3, you were asked to figure out what options the cell might reasonably have to reoxidize the NADH to NAD+ in order to avoid consuming the available pools of NAD+ and to thus avoid stopping glycolysis. Put differently, during glycolysis, cells can generate large amounts of NADH and slowly exhaust their supplies of NAD+. If glycolysis is to continue, the cell must find a way to regenerate NAD+, either by synthesis or by some form of recycling.

    In the absence of any other process—that is, if we consider glycolysis alone—it is not immediately obvious what the cell might do. One choice is to try putting the electrons that were once stripped off of the glucose derivatives right back onto the downstream product, pyruvate, or one of its derivatives. We can generalize the process by describing it as the returning of electrons to the molecule that they were once removed, usually to restore pools of an oxidizing agent. This, in short, is fermentation. As we will discuss in a different section, the process of respiration can also regenerate the pools of NAD+ from NADH. Cells lacking respiratory chains or in conditions where using the respiratory chain is unfavorable may choose fermentation as an alternative mechanism for garnering energy from small molecules.

    An example: lactic acid fermentation

    An everyday example of a fermentation reaction is the reduction of pyruvate to lactate by the lactic acid fermentation reaction. This reaction should be familiar to you: it occurs in our muscles when we exert ourselves during exercise. When we exert ourselves, our muscles require large amounts of ATP to perform the work we are demanding of them. As the ATP is consumed, the muscle cells are unable to keep up with the demand for respiration, O2 becomes limiting, and NADH accumulates. Cells need to get rid of the excess and regenerate NAD+, so pyruvate serves as an electron acceptor, generating lactate and oxidizing NADH to NAD+. Many bacteria use this pathway as a way to complete the NADH/NAD+ cycle. You may be familiar with this process from products like sauerkraut and yogurt. The chemical reaction of lactic acid fermentation is the following:

    Pyruvate + NADH ↔ lactic acid + NAD+

    lactate_fermentation2.png

    Figure 1. Lactic acid fermentation converts pyruvate (a slightly oxidized carbon compound) to lactic acid. In the process, NADH is oxidized to form NAD+. Attribution: Marc T. Facciotti (original work)

    Energy story for the fermentation of pyruvate to lactate

    An example (if a bit lengthy) energy story for lactic acid fermentation is the following:

    The reactants are pyruvate, NADH, and a proton. The products are lactate and NAD+. The process of fermentation results in the reduction of pyruvate to form lactic acid and the oxidation of NADH to form NAD+. Electrons from NADH and a proton are used to reduce pyruvate into lactate. If we examine a table of standard reduction potential, we see under standard conditions that a transfer of electrons from NADH to pyruvate to form lactate is exergonic and thus thermodynamically spontaneous. The reduction and oxidation steps of the reaction are coupled and catalyzed by the enzyme lactate dehydrogenase.

    A second example: alcohol fermentation

    Another familiar fermentation process is alcohol fermentation, which produces ethanol, an alcohol. The alcohol fermentation reaction is the following:

    alcohol_fermentation.png

    Figure 2. Ethanol fermentation is a two-step process. Pyruvate (pyruvic acid) is first converted into carbon dioxide and acetaldehyde. The second step converts acetaldehyde to ethanol and oxidizes NADH to NAD+. Attribution: Marc T. Facciotti (original work)

    In the first reaction, a carboxyl group is removed from pyruvic acid, releasing carbon dioxide as a gas (some of you may be familiar with this as a key component of various beverages). The second reaction removes electrons from NADH, forming NAD+ and producing ethanol (another familiar compound—usually in the same beverage) from the acetaldehyde, which accepts the electrons.

    Suggested discussion

    Write a complete energy story for alcohol fermentation. Propose possible benefits of this type of fermentation for the single-celled yeast organism.

    Potentially helpful videos

    Here is a ChemWiki link on fermentation reactions.

    Fermentation pathways are numerous

    While the lactic acid fermentation and alcohol fermentation pathways described above are examples, there are many more reactions (too numerous to go over) that Nature has evolved to complete the NADH/NAD+ cycle. It is important that you understand the general concepts behind these reactions. In general, cells try to maintain a balance or constant ratio between NADH and NAD+; when this ratio becomes unbalanced, the cell compensates by modulating other reactions to compensate. The only requirement for a fermentation reaction is that it uses a small organic compound as an electron acceptor for NADH and regenerates NAD+. Other familiar fermentation reactions include ethanol fermentation (as in beer and bread), propionic fermentation (it's what makes the holes in Swiss cheese), and malolactic fermentation (it's what gives Chardonnay its more mellow flavor—the more conversion of malate to lactate, the softer the wine). In Figure 3, you can see a large variety of fermentation reactions that various bacteria use to reoxidize NADH to NAD+. All of these reactions start with pyruvate or a derivative of pyruvate metabolism, such as oxaloacetate or formate. Pyruvate is produced from the oxidation of sugars (glucose or ribose) or other small, reduced organic molecules. It should also be noted that other compounds can be used as fermentation substrates besides pyruvate and its derivatives. These include methane fermentation, sulfide fermentation, or the fermentation of nitrogenous compounds such as amino acids. You are not expected to memorize all of these pathways. You are, however, expected to recognize a pathway that returns electrons to products of the compounds that were originally oxidized to recycle the NAD+/NADH pool and to associate that process with fermentation.

    Figure 3: Various fermentation pathways using pyruvate as the initial substrate. In the figure, pyruvate is reduced to a variety of products via different and sometimes multistep (dashed arrows represent possible multiple step processes) reactions. All details are deliberately not shown. The key point is to appreciate that fermentation is a broad term not solely associated with the conversion of pyruvate to lactic acid or ethanol.   Source: Original work Marc T. Facciotti

    Figure 3. This figure shows various fermentation pathways using pyruvate as the initial substrate. In the figure, pyruvate is reduced to a variety of products via different and sometimes multistep (dashed arrows represent possible multistep processes) reactions. All details are deliberately not shown. The key point is to appreciate that fermentation is a broad term not solely associated with the conversion of pyruvate to lactic acid or ethanol. Source: Marc T. Facciotti (original work)

    A note on the link between substrate-level phosphorylation and fermentation

    Fermentation occurs in the absence of molecular oxygen (O2). It is an anaerobic process. Notice there is no O2 in any of the fermentation reactions shown above. Many of these reactions are quite ancient, hypothesized to be some of the first energy-generating metabolic reactions to evolve. This makes sense if we consider the following:

    1. The early atmosphere was highly reduced, with little molecular oxygen readily available.
    2. Small, highly reduced organic molecules were relatively available, arising from a variety of chemical reactions.
    3. These types of reactions, pathways, and enzymes are found in many different types of organisms, including bacteria, archaea, and eukaryotes, suggesting these are very ancient reactions.
    4. The process evolved long before O2 was found in the environment.
    5. The substrates, highly reduced, small organic molecules, like glucose, were readily available.
    6. The end products of many fermentation reactions are small organic acids, produced by the oxidation of the initial substrate.
    7. The process is coupled to substrate-level phosphorylation reactions. That is, small, reduced organic molecules are oxidized, and ATP is generated by first a redox reaction followed by the substrate-level phosphorylation.
    8. This suggests that substrate-level phosphorylation and fermentation reactions coevolved.

    Suggested discussion

    If the hypothesis is correct that substrate-level phosphorylation and fermentation reactions coevolved and were the first forms of energy metabolism that cells used to generate ATP, then what would be the consequences of such reactions over time? What if these were the only forms of energy metabolism available over hundreds of thousands of years? What if cells were isolated in a small, closed environment? What if the small, reduced substrates were not being produced at the same rate of consumption during this time?

    Consequences of fermentation

    Imagine a world where fermentation is the primary mode for extracting energy from small molecules. As populations thrive, they reproduce and consume the abundance of small, reduced organic molecules in the environment, producing acids. One consequence is the acidification (decrease in pH) of the environment, including the internal cellular environment. This is not so good, since changes in pH can have a profound influence on the function and interactions among various biomolecules. Therefore, mechanisms needed to evolved that could remove the various acids. Fortunately, in an environment rich in reduced compounds, substrate-level phosphorylation and fermentation can produce large quantities of ATP.

    It is hypothesized that this scenario was the beginning of the evolution of the F0F1-ATPase, a molecular machine that hydrolyzes ATP and translocates protons across the membrane (we'll see this again in the next section). With the F0F1-ATPase, the ATP produced from fermentation could now allow for the cell to maintain pH homeostasis by coupling the free energy of hydrolysis of ATP to the transport of protons out of the cell. The downside is that now cells are pumping all of these protons into the environment, which will now start to acidify.

    Suggested discussion

    If the hypothesis is correct that the F0F1-ATPase also coevolved with substrate-level phosphorylation and fermentation reactions, then what would happen over time to the environment? While small, reduced organic compounds may have been initially abundant, if fermentation "took off" at some point, then the reduced compounds would run out and ATP might then become scarce as well. That's a problem. Thinking with the design challenge rubric in mind, define the problem(s) facing the cell in this hypothesized environment. What are other potential mechanisms or ways Nature could overcome the problem(s)?


    Lecture 08: Redox reactions is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.

    • Was this article helpful?