Buffer and Extraction Buffer- Definition, Components, Significance
Updated:
An extraction buffer is a type of buffer solution that is used to break open cells and release their contents for further analysis. Buffer solutions are mixtures of weak acids and their conjugate bases, or weak bases and their conjugate acids, that resist changes in pH when small amounts of acid or base are added. By maintaining a constant pH, buffer solutions help to preserve the structure and function of biological molecules, such as proteins and nucleic acids, that are sensitive to changes in acidity or alkalinity.
Extraction buffers are designed to lyse (break open) cells by disrupting their membranes and solubilizing their components. Lysis can be achieved by physical methods, such as grinding, sonication, or freeze-thawing, or by chemical methods, such as detergents, chaotropic agents, or enzymes. The choice of lysis method depends on the type of cell and the target molecule of interest. For example, bacterial cells have a tough cell wall that requires mechanical or enzymatic lysis, while animal cells have a more fragile membrane that can be lysed by mild detergents.
The composition of an extraction buffer depends on the purpose of the experiment and the properties of the target molecule. Different extraction buffers may contain different salts, pH adjusters, reducing agents, enzyme inhibitors, substrates, cofactors, chelators, stabilizers, or preservatives. These components help to optimize the yield and quality of the extracted molecules by preventing degradation, aggregation, oxidation, precipitation, or contamination. Some examples of common extraction buffers are:
- RIPA buffer: A detergent-based buffer that lyses most types of cells and solubilizes proteins and nucleic acids. It contains sodium chloride (NaCl), tris(hydroxymethyl)aminomethane (Tris), sodium deoxycholate (SDC), sodium dodecyl sulfate (SDS), ethylenediaminetetraacetic acid (EDTA), and phenylmethylsulfonyl fluoride (PMSF).
- Tris-EDTA (TE) buffer: A mild buffer that preserves nucleic acids by chelating metal ions that can catalyze nuclease activity. It contains Tris and EDTA.
- Phosphate-buffered saline (PBS): A physiological buffer that mimics the osmolarity and pH of biological fluids. It contains sodium phosphate (Na2HPO4), sodium chloride (NaCl), and potassium chloride (KCl).
- Tris-buffered saline (TBS): A similar buffer to PBS but with Tris instead of phosphate. It is often used for immunological assays.
- Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) sample buffer: A buffer that denatures proteins and coats them with negative charges for electrophoretic separation. It contains Tris, SDS, glycerol, bromophenol blue (a dye), and dithiothreitol (DTT) or b-mercaptoethanol (BME) as reducing agents.
In summary, an extraction buffer is a buffer solution that is used to break open cells and release their contents for further analysis. It helps to maintain a constant pH and to protect the extracted molecules from degradation or loss of activity. Different extraction buffers have different compositions depending on the type of cell and the target molecule of interest.
An extraction buffer is a solution that contains various components that help to break open the cells and stabilize the compounds of interest. The components of an extraction buffer may vary depending on the type of cells and compounds being extracted, but some common components are:
-
Buffering agents: These are substances that maintain the pH of the solution within a certain range. The pH of the solution affects the charge and solubility of the compounds, as well as the activity and stability of enzymes. Buffering agents also prevent drastic changes in pH that may occur due to cellular metabolism or environmental factors. Some examples of buffering agents are Tris, phosphate, HEPES, and MOPS.
-
Salts: These are substances that regulate the osmolarity and ionic strength of the solution. The osmolarity and ionic strength affect the hydration and interactions of the compounds, as well as the permeability and integrity of the cell membranes. Salts also help to balance the charge of the compounds and prevent aggregation or precipitation. Some examples of salts are sodium chloride, potassium chloride, magnesium chloride, and calcium chloride.
-
Detergents: These are substances that disrupt the lipid bilayer of the cell membranes and solubilize the membrane-bound proteins and lipids. Detergents also help to reduce the surface tension and viscosity of the solution, facilitating the mixing and extraction process. Some examples of detergents are sodium dodecyl sulfate (SDS), Triton X-100, Nonidet P-40 (NP-40), and Tween 20.
-
Reducing agents: These are substances that prevent or reverse the oxidation of thiol groups in proteins. Thiol groups are present in cysteine residues and can form disulfide bonds within or between proteins. Disulfide bonds can affect the structure and function of proteins, as well as their solubility and stability. Reducing agents also scavenge free radicals and reactive oxygen species that may damage the compounds. Some examples of reducing agents are dithiothreitol (DTT), beta-mercaptoethanol (BME), cysteine, and glutathione.
-
Protease inhibitors: These are substances that inhibit or block the activity of proteases, which are enzymes that degrade proteins. Proteases can be released from lysed cells or from contaminating microorganisms, and can cause unwanted proteolysis of the compounds of interest. Protease inhibitors also protect the proteins from autolysis or self-digestion. Some examples of protease inhibitors are phenylmethylsulfonyl fluoride (PMSF), iodoacetamide, pepstatin A, leupeptin, and aprotinin.
-
Other additives: These are substances that have specific roles or functions depending on the type of cells and compounds being extracted. Some examples of other additives are:
- Substrates and cofactors: These are molecules that bind to enzymes and modulate their activity and stability. Substrates are molecules that are acted upon by enzymes, while cofactors are molecules that assist enzymes in catalysis. Substrates and cofactors can be added to extraction buffers to maintain or enhance enzyme activity during extraction.
- EDTA: This is a chelating agent that binds to divalent metal ions such as calcium, magnesium, zinc, and iron. Metal ions can react with thiol groups in proteins and form mercaptides, which can alter protein structure and function. Metal ions can also activate metalloproteases, which are proteases that require metal ions for catalysis. EDTA can be added to extraction buffers to remove metal ions and prevent these effects.
- Polyvinylpyrrolidone (PVP): This is a polymer that binds to phenolic compounds in plant tissues. Phenolic compounds can interact with proteins by non-covalent forces such as hydrophobic, ionic, and hydrogen bonds, causing protein precipitation. Phenolic compounds can also be oxidized by phenol oxidases to form quinones, which can react with proteins and form covalent bonds. PVP can be added to extraction buffers to adsorb phenolic compounds and prevent these effects.
- Sodium azide: This is an antibacterial agent that inhibits cytochrome c oxidase in bacteria. Cytochrome c oxidase is an enzyme that is involved in cellular respiration and energy production. Sodium azide can be added to extraction buffers to prevent bacterial growth and contamination.
These are some of the common components of an extraction buffer. However, it is important to note that different extraction buffers may have different compositions and concentrations depending on the specific needs and preferences of the experiment. Therefore, it is advisable to optimize the extraction buffer for each experiment and test its efficacy and compatibility with the compounds of interest.
An anti-oxidant is a substance that prevents or slows down the oxidation of other molecules. Oxidation is a chemical reaction that involves the loss of electrons or the increase of oxidation state. Oxidation can cause damage to the structure and function of biomolecules, such as proteins, lipids, and nucleic acids.
One of the main sources of oxidative stress in biological systems is the reactive oxygen species (ROS), which are highly reactive molecules that contain oxygen. ROS can be generated by various metabolic processes, environmental factors, or cellular stress. ROS can react with biomolecules and cause oxidative modifications, such as peroxidation, carbonylation, nitration, and cross-linking. These modifications can alter the conformation, activity, stability, and interactions of biomolecules.
To protect biomolecules from oxidative damage, cells have developed various antioxidant systems that scavenge ROS and repair oxidative modifications. These systems include enzymatic antioxidants, such as superoxide dismutase, catalase, and glutathione peroxidase; and non-enzymatic antioxidants, such as glutathione, ascorbic acid, tocopherol, and uric acid.
However, when cells are lysed for extraction purposes, the antioxidant systems may be disrupted or overwhelmed by the sudden increase of ROS. Therefore, it is important to add exogenous antioxidants to the extraction buffer to maintain a reducing environment and prevent oxidative damage to the extracted biomolecules.
Some of the commonly used antioxidants in extraction buffers are:
- Dithiothreitol (DTT): This is a small molecule that contains two thiol groups that can reduce disulfide bonds in proteins and other biomolecules. DTT can also scavenge ROS and protect thiol groups from oxidation.
- Beta-mercaptoethanol (BME): This is another small molecule that contains one thiol group that can reduce disulfide bonds and scavenge ROS. BME is more volatile and toxic than DTT, but it is cheaper and more stable.
- Cysteine: This is an amino acid that contains a thiol group that can reduce disulfide bonds and scavenge ROS. Cysteine can also act as a precursor for glutathione synthesis in cells.
- Reduced glutathione (GSH): This is a tripeptide that contains a thiol group that can reduce disulfide bonds and scavenge ROS. GSH can also participate in various enzymatic reactions that involve redox regulation.
The choice of antioxidant depends on several factors, such as the type of biomolecule to be extracted, the stability and solubility of the antioxidant, the compatibility with other reagents in the buffer, and the cost and availability of the antioxidant.
The concentration of antioxidant should be optimized according to the experimental conditions and the level of oxidative stress. Generally, higher concentrations of antioxidant are required for extracting proteins than nucleic acids or lipids. The concentration of antioxidant should also be balanced with the concentration of other reagents in the buffer, such as salts, detergents, chelators, and inhibitors.
The role of an antioxidant in an extraction buffer is to preserve the integrity and functionality of the extracted biomolecules by preventing or minimizing oxidative damage caused by ROS or other oxidizing agents. By adding an antioxidant to the extraction buffer, one can improve the yield and quality of the extraction process and enhance the reliability and reproducibility of the downstream analysis.
Once the cell is disrupted, proteolytic enzymes that were carefully packaged and controlled within the intact cells are released which may start to degrade proteins in the extract, including the protein of interest. To slow down unwanted proteolysis, all extraction and purification steps are carried out at 4°C, and in addition a range of protease inhibitors is included in the buffer. Common examples of inhibitors include:
- Di-isopropylphosphofluoridate (DFP), phenylmethyl sulphonylfluoride (PMSF) and tosylphenylalanyl-chloromethylketone (TPCK): These are all serine protease inhibitors that bind covalently to the active site serine residue of enzymes such as trypsin, chymotrypsin and elastase. They are effective at low concentrations (0.1-1 mM) but they are also toxic and unstable in aqueous solutions, so they should be added just before use.
- Iodoacetate and cystatin: These are thiol protease inhibitors that react with the cysteine residue in the active site of enzymes such as papain and cathepsins. Iodoacetate is irreversible and cystatin is reversible. They are used at concentrations of 1-10 mM and 0.1-1 mg/ml, respectively.
- Pepstatin: This is an aspartic protease inhibitor that binds non-covalently to the active site of enzymes such as pepsin and renin. It is used at concentrations of 1-10 µM.
- EDTA and 1,10-phenanthroline: These are metalloprotease inhibitors that chelate the metal ions (such as Zn2+, Mg2+, Ca2+, etc.) required for the catalytic activity of enzymes such as thermolysin and carboxypeptidase. They are used at concentrations of 1-10 mM and 0.1-1 mM, respectively.
- Amastatin and bestatin: These are exopeptidase inhibitors that inhibit the removal of amino acids from the N- or C-terminal ends of peptides by enzymes such as aminopeptidase and dipeptidyl peptidase. They are used at concentrations of 0.1-1 mM and 10-100 µM, respectively.
The choice and concentration of enzyme inhibitors depend on the type and source of the cells, the protein of interest and the downstream applications. It is advisable to use a cocktail of inhibitors that covers a broad spectrum of proteases to ensure maximum protection of the proteins in the extract.
Enzymes are biological catalysts that speed up biochemical reactions in living organisms and can be extracted from cells for various purposes. However, enzymes are often sensitive to changes in temperature, pH, ionic strength and other factors that can affect their structure and function. Therefore, extraction buffers need to contain some components that can help stabilize and maintain the activity of the enzymes during the extraction and purification processes.
One of these components is the enzyme substrate, which is the molecule that binds to the active site of the enzyme and undergoes a chemical reaction. The binding of substrate to the enzyme is very specific and often involves a conformational change in the enzyme that enhances its catalytic efficiency. This binding also protects the enzyme from denaturation, aggregation or degradation by other molecules or agents .
Low levels of substrate are often included in extraction buffers when an enzyme is purified, since binding of substrate to the enzyme active site can stabilise the enzyme during purification processes. For example, glucose-6-phosphate dehydrogenase, an enzyme involved in carbohydrate metabolism, can be stabilized by adding glucose-6-phosphate as a substrate to the extraction buffer.
Another component that can affect the enzyme activity is the cofactor, which is a non-protein molecule or ion that is required for the enzyme to function properly. Cofactors can be either organic molecules, such as vitamins or coenzymes, or inorganic ions, such as metal ions. Cofactors can bind to the enzyme either tightly (prosthetic groups) or loosely (co-substrates) and can participate in the catalytic reaction by transferring electrons, atoms or groups of atoms.
Some cofactors are essential for the enzyme activity and cannot be replaced by other molecules. For example, zinc is a cofactor for carbonic anhydrase, an enzyme that catalyzes the hydration of carbon dioxide. Without zinc, carbonic anhydrase cannot function at all. Other cofactors are more flexible and can be substituted by similar molecules. For example, nicotinamide adenine dinucleotide (NAD) and nicotinamide adenine dinucleotide phosphate (NADP) are two coenzymes that can act as electron acceptors for different enzymes.
Where relevant, cofactors that otherwise might be lost during purification are also included in the extraction buffer to maintain enzyme activity. For example, magnesium ions are often added to extraction buffers for enzymes that use adenosine triphosphate (ATP) as a substrate or co-substrate, such as kinases or ATPases.
Therefore, enzyme substrate and cofactors play an important role in stabilizing and preserving the activity of enzymes during extraction and purification processes. By adding appropriate amounts and types of these components to the extraction buffer, the quality and yield of the enzyme product can be improved.
EDTA stands for ethylenediaminetetraacetic acid, a chelating agent that can bind to metal ions and form stable complexes. EDTA is often added to extraction buffers to remove divalent metal ions such as calcium, magnesium, iron, zinc and copper from the cell lysate. These metal ions can interfere with the extraction and purification of proteins and nucleic acids in several ways:
- They can react with thiol groups in proteins and form mercaptides, which can alter the protein structure and function.
- They can activate metalloproteases, which are enzymes that degrade proteins by cleaving peptide bonds at specific sites.
- They can catalyze the oxidation of phenolic compounds in plant tissues, leading to the formation of quinones that can precipitate proteins.
- They can inhibit the activity of some enzymes that require metal cofactors, such as DNA polymerase and reverse transcriptase.
By chelating the metal ions, EDTA prevents these unwanted reactions and preserves the integrity and activity of the biomolecules of interest. EDTA also helps to maintain the pH of the buffer by acting as a weak acid. However, EDTA can also have some drawbacks in certain applications:
- It can interfere with some downstream assays that rely on metal ions, such as colorimetric assays and PCR.
- It can reduce the solubility of some proteins that depend on metal ions for their stability, such as metalloenzymes and metal-binding proteins.
- It can affect the binding affinity of some antibodies that recognize metal-dependent epitopes on antigens.
Therefore, the concentration and type of EDTA used in an extraction buffer should be carefully optimized according to the specific purpose and target molecule of the experiment.
Polyvinylpyrrolidone (PVP) is a synthetic polymer that has many applications in medicine, cosmetics, food and biotechnology. One of its uses is to enhance the extraction of proteins from plant tissues. Plant tissues contain considerable amounts of phenolic compounds that can bind to enzymes and other proteins by non-covalent forces, including hydrophobic, ionic and hydrogen bonds, causing protein precipitation. This can reduce the yield and quality of the extracted proteins and interfere with their downstream analysis. Insoluble PVP (which mimics the polypeptide backbone) is therefore added to extraction buffers for plant tissue to adsorb the phenolic compounds which can then be removed by centrifugation. PVP acts as a molecular sponge that sequesters the phenolic compounds and prevents them from interacting with the proteins of interest. PVP also helps to stabilize the proteins by reducing their aggregation and denaturation.
The concentration of PVP in the extraction buffer depends on the type and amount of plant tissue and the phenolic content. Typically, PVP concentrations range from 1% to 10% (w/v). The molecular weight of PVP also affects its performance. Higher molecular weight PVP (such as PVP-40) has a higher binding capacity for phenolic compounds than lower molecular weight PVP (such as PVP-10). However, higher molecular weight PVP may also bind to some proteins and reduce their solubility. Therefore, a balance between the binding capacity and the solubility of PVP should be considered when choosing the optimal PVP for a specific extraction buffer.
In addition to PVP, thiol compounds (reducing agents) are also added to extraction buffers for plant tissue to minimise the activity of phenol oxidases, which are enzymes that catalyse the oxidation of phenolic compounds to quinones. Quinones are highly reactive molecules that can form covalent bonds with proteins and cause irreversible damage. Thiol compounds such as dithiothreitol (DTT), b-mercaptoethanol (BME), cysteine or reduced glutathione can scavenge quinones and protect the proteins from oxidation.
PVP is an important component of extraction buffers for plant tissue as it helps to remove phenolic compounds that can interfere with protein extraction and analysis. By using PVP in combination with thiol compounds, the yield and quality of extracted proteins from plant tissue can be improved.
Sodium azide is a chemical compound that has the formula NaN3. It is a colorless, odorless, and highly soluble salt that decomposes rapidly when heated or exposed to acids. Sodium azide is often used as a preservative in extraction buffers because it inhibits the growth of bacteria and fungi that can contaminate the samples and degrade the proteins of interest. Sodium azide acts as a bacteriostatic agent by interfering with the electron transport chain in the bacterial cell membrane, thus preventing the synthesis of ATP and causing cell death. Sodium azide can also inhibit some enzymes that require metal ions as cofactors, such as cytochrome oxidase and catalase.
Sodium azide is usually added to extraction buffers at low concentrations, typically 0.02% to 0.1%. However, sodium azide can be toxic to humans and animals if ingested, inhaled, or absorbed through the skin. It can cause symptoms such as nausea, vomiting, headache, hypotension, tachycardia, dyspnea, cyanosis, and convulsions. In severe cases, sodium azide can cause cardiac arrest and death. Therefore, sodium azide should be handled with care and disposed of properly. Sodium azide should not be mixed with acids or metals, as this can generate explosive or toxic gases. Sodium azide should also be avoided in buffers that are used for enzymatic assays or immunological detection methods that involve metal-containing reagents.
Sodium azide is an effective and widely used preservative in extraction buffers, but it has some limitations and risks that should be considered. Sodium azide can help prevent bacterial and fungal contamination in the samples and protect the proteins of interest from degradation. However, sodium azide can also inhibit some enzymes and interfere with some analytical methods that require metal ions. Moreover, sodium azide can be hazardous to human health and the environment if not handled and disposed of properly. Therefore, sodium azide should be used with caution and only when necessary in extraction buffers.
The primary purpose of an extraction buffer is to isolate the compounds of interest and keep them in a stable environment. They are of enormous importance in practical biochemical work as many biochemical molecules are weak electrolytes and their ionic status varies with pH. Hence, there is a need to stabilise this ionic status during the course of a practical experiment .
An extraction buffer also helps in the lysis of the cell wall and nuclear membrane, releasing the nucleic acids from the cell . The buffer should be compatible with the target constituent of the cell and should not interfere with its structure or function. For example, Tris buffer is commonly used for DNA extraction as it has a pH range that is suitable for DNA stability .
An extraction buffer may also contain various additives that protect the nucleic acids from degradation, oxidation, or contamination. These include reducing agents, enzyme inhibitors, substrate and cofactors, EDTA, PVP, and sodium azide . These additives can enhance the yield and quality of the nucleic acids extracted and prevent unwanted reactions that may compromise the downstream applications.
An extraction buffer is therefore a crucial component of any molecular biology experiment that involves nucleic acid analysis. It enables the efficient and reliable extraction of nucleic acids from various sources and preserves their integrity for further processing and detection. An extraction buffer can also be tailored to suit different types of nucleic acids, cells, and experimental conditions.
We are Compiling this Section. Thanks for your understanding.