U.S. PHARMACOPEIA

Search USP29  
1047 BIOTECHNOLOGY-DERIVED ARTICLES—TESTS
The emergence of drug macromolecules obtained through biotechnological processes has led to a set of specialized tests and assays to determine quality, identity, purity, and potency of these articles in addition to the methods traditionally used for other drug products. These specialized tests that are presented below are Amino Acid Analysis, Capillary Electrophoresis, Isoelectric Focusing, Peptide Mapping, Polyacrylamide Gel Electrophoresis, and Total Protein Assay.

AMINO ACID ANALYSIS
Amino acid analysis refers to the methodology used to determine the amino acid composition or content of proteins, peptides, and other pharmaceutical preparations. Proteins and peptides are macromolecules consisting of covalently bonded amino acid residues organized as a linear polymer. The sequence of the amino acids in a protein or peptide determines the properties of the molecule. Proteins are considered large molecules that commonly exist as folded structures with a specific conformation, while peptides are smaller and may consist of only a few amino acids. Amino acid analysis can be used to quantify protein and peptides, to determine the identity of proteins or peptides based on their amino acid composition, to support protein and peptide structure analysis, to evaluate fragmentation strategies for peptide mapping, and to detect atypical amino acids that might be present in a protein or peptide. It is necessary to hydrolyze a protein/peptide to its individual amino acid constituents before amino acid analysis. Following protein/peptide hydrolysis, the amino acid analysis procedure can be the same as that practiced for free amino acids in other pharmaceutical preparations. The amino acid constituents of the test sample are typically derivatized for analysis.
Apparatus
Methods used for amino acid analysis are usually based on a chromatographic separation of the amino acids present in the test sample. Current techniques take advantage of the automated chromatographic instrumentation designed for analytical methodologies. An amino acid analysis instrument will typically be a low-pressure or high-pressure liquid chromatograph capable of generating mobile phase gradients that separate the amino acid analytes on a chromatographic column. The instrument must have postcolumn derivatization capability, unless the sample is analyzed using precolumn derivatization. The detector is usually a UV-visible or fluorescence detector depending on the derivatization method used. A recording device (e.g., integrator) is used for transforming the analog signal from the detector and for quantitation. It is preferred that instrumentation be dedicated particularly for amino acid analysis.
General Precautions
Background contamination is always a concern for the analyst in performing amino acid analysis. High-purity reagents are necessary (e.g., low-purity hydrochloric acid can contribute to glycine contamination). Analytical reagents are changed routinely every few weeks using only high-pressure liquid chromatography (HPLC) grade solvents. Potential microbial contamination and foreign material that might be present in the solvents are reduced by filtering solvents before use, keeping solvent reservoirs covered, and not placing amino acid analysis instrumentation in direct sunlight.
Laboratory practices can determine the quality of the amino acid analysis. Place the instrumentation in a low traffic area of the laboratory. Keep the laboratory clean. Clean and calibrate pipets according to a maintenance schedule. Keep pipet tips in a covered box; the analysts may not handle pipet tips with their hands. The analysts may wear powder-free latex or equivalent gloves. Limit the number of times a test sample vial is opened and closed because dust can contribute to elevated levels of glycine, serine, and alanine.
A well-maintained instrument is necessary for acceptable amino acid analysis results. If the instrument is used on a routine basis, it is to be checked daily for leaks, detector and lamp stability, and the ability of the column to maintain resolution of the individual amino acids. Clean or replace all instrument filters and other maintenance items on a routine schedule.
Reference Standard Material
Acceptable amino acid standards are commercially available1 for amino acid analysis and typically consist of an aqueous mixture of amino acids. When determining amino acid composition, protein or peptide standards are analyzed with the test material as a control to demonstrate the integrity of the entire procedure. Highly purified bovine serum albumin has been used as a protein standard for this purpose.
Calibration of Instrumentation
Calibration of amino acid analysis instrumentation typically involves analyzing the amino acid standard, which consists of a mixture of amino acids at a number of concentrations, to determine the response factor and range of analysis for each amino acid. The concentration of each amino acid in the standard is known. In the calibration procedure, the analyst dilutes the amino acid standard to several different analyte levels within the expected linear range of the amino acid analysis technique. Then, replicates at each of the different analyte levels can be analyzed. Peak areas obtained for each amino acid are plotted versus the known concentration for each of the amino acids in the standard dilution. These results will allow the analyst to determine the range of amino acid concentrations where the peak area of a given amino acid is an approximately linear function of the amino acid concentration. It is important that the analyst prepare the samples for amino acid analysis so that they are within the analytical limits (e.g., linear working range) of the technique employed in order to obtain accurate and repeatable results.
Four to six amino acid standard levels are analyzed to determine a response factor for each amino acid. The response factor is calculated as the average peak area or peak height per nmol of amino acid present in the standard. A calibration file consisting of the response factor for each amino acid is prepared and is used to calculate the concentration of each amino acid present in the test sample. This calculation involves dividing the peak area corresponding to a given amino acid by the response factor for that amino acid to give the nmol of the amino acid. For routine analysis, a single-point calibration may be sufficient; however, the calibration file is updated frequently and tested by the analysis of analytical controls to ensure its integrity.
Repeatability
Consistent high quality amino acid analysis results from an analytical laboratory require attention to the repeatability of the assay. During analysis of the chromatographic separation of the amino acids or their derivatives, numerous peaks can be observed on the chromatogram that corresponds to the amino acids. The large number of peaks makes it necessary to have an amino acid analysis system that can repeatedly identify the peaks based on retention time and integrate the peak areas for quantitation. A typical repeatability evaluation involves preparing a standard amino acid solution and analyzing many replicates (i.e., six analyses or more) of the same standard solution. The relative standard deviation (RSD) is determined for the retention time and integrated peak area of each amino acid. An evaluation of the repeatability is expanded to include multiple assays conducted over several days by different analysts. Multiple assays include the preparation of standard dilutions from starting materials to determine the variation due to sample handling. Often, the amino acid composition of a standard protein (e.g., bovine serum albumin) is analyzed as part of the repeatability evaluation. By evaluating the replicate variation (i.e., RSD), the laboratory can establish analytical limits to ensure that the analyses from the laboratory are under control. It is desirable to establish the lowest practical variation limits to ensure the best results. Areas to focus on to lower the variability of the amino acid analysis include sample preparation, high background spectral interference due to quality of reagents and/or laboratory practices, instrument performance and maintenance, data analysis and interpretation, and analyst performance and habits. All parameters involved are fully investigated in the scope of the validation work.
Sample Preparation
Accurate results from amino acid analysis require purified protein and peptide samples. Buffer components (e.g., salts, urea, detergents) can interfere with the amino acid analysis and are removed from the sample before analysis. Methods that utilize postcolumn derivatization of the amino acids are generally not affected by buffer components to the extent seen with precolumn derivatization methods. It is desirable to limit the number of sample manipulations to reduce potential background contamination, to improve analyte recovery, and to reduce labor. Common techniques used to remove buffer components from protein samples include the following methods: (1) injecting the protein sample onto a reverse-phase HPLC system, removing the protein with a volatile solvent containing a sufficient organic component, and drying the sample in a vacuum centrifuge; (2) dialysis against a volatile buffer or water; (3) centrifugal ultrafiltration for buffer replacement with a volatile buffer or water; (4) precipitating the protein from the buffer using an organic solvent (e.g., acetone); and (5) gel filtration.
Internal Standards
It is recommended that an internal standard be used to monitor physical and chemical losses and variations during amino acid analysis. An accurately known amount of internal standard can be added to a protein solution prior to hydrolysis. The recovery of the internal standard gives the general recovery of the amino acids from the protein solution. Free amino acids, however, do not behave in the same way as protein-bound amino acids during hydrolysis because their rates of release or destruction are variable. Therefore, the use of an internal standard to correct for losses during hydrolysis may give unreliable results. It will be necessary to take this particular point into consideration when interpreting the results. Internal standards can also be added to the mixture of amino acids after hydrolysis to correct for differences in sample application and changes in reagent stability and flow rates. Ideally, an internal standard is an unnaturally occurring primary amino acid that is commercially available and inexpensive. It should also be stable during hydrolysis, its response factor should be linear with concentration, and it needs to elute with a unique retention time without overlapping other amino acids. Commonly used amino acid standards include norleucine, nitrotyrosine, and -aminobutyric acid.
Protein Hydrolysis
Hydrolysis of protein and peptide samples is necessary for amino acid analysis of these molecules. The glassware used for hydrolysis must be very clean to avoid erroneous results. Glove powders and fingerprints on hydrolysis tubes may cause contamination. To clean glass hydrolysis tubes, boil tubes for 1 hour in 1 N hydrochloric acid or soak tubes in concentrated nitric acid or in a mixture of concentrated hydrochloric acid and concentrated nitric acid (1:1). Clean hydrolysis tubes are rinsed with high-purity water followed by a rinse with HPLC grade methanol, dried overnight in an oven, and stored covered until use. Alternatively, pyrolysis of clean glassware at 500 for 4 hours may also be used to eliminate contamination from hydrolysis tubes. Adequate disposable laboratory material can also be used.
Acid hydrolysis is the most common method for hydrolyzing a protein sample before amino acid analysis. The acid hydrolysis technique can contribute to the variation of the analysis due to complete or partial destruction of several amino acids. Tryptophan is destroyed; serine and threonine are partially destroyed; methionine might undergo oxidation; and cysteine is typically recovered as cystine (but cystine recovery is usually poor because of partial destruction or reduction to cysteine). Application of adequate vacuum (less than 200 µm of mercury or 26.7 Pa) or introduction of an inert gas (argon) in the headspace of the reaction vessel can reduce the level of oxidative destruction. In peptide bonds involving isoleucine and valine, the amido bonds of Ile-Ile, Val-Val, Ile-Val, and Val-Ile are partially cleaved; and asparagine and glutamine are deamidated, resulting in aspartic acid and glutamic acid, respectively. The loss of tryptophan, asparagine, and glutamine during an acid hydrolysis limits quantitation to 17 amino acids. Some of the hydrolysis techniques described are used to address these concerns. Some of the hydrolysis techniques described (i.e., Methods 4–11) may cause modifications to other amino acids. Therefore, the benefits of using a given hydrolysis technique are weighed against the concerns with the technique and are tested adequately before employing a method other than acid hydrolysis.
A time-course study (i.e., amino acid analysis at acid hydrolysis times of 24, 48, and 72 hours) is often employed to analyze the starting concentration of amino acids that are partially destroyed or slow to cleave. By plotting the observed concentration of labile amino acids (i.e., serine and threonine) versus hydrolysis time, the line can be extrapolated to the origin to determine the starting concentration of these amino acids. Time-course hydrolysis studies are also used with amino acids that are slow to cleave (e.g., isoleucine and valine). During the hydrolysis time course, the analyst will observe a plateau in these residues. The level of this plateau is taken as the residue concentration. If the hydrolysis time is too long, the residue concentration of the sample will begin to decrease, indicating destruction by the hydrolysis conditions.
An acceptable alternative to the time-course study is to subject an amino acid calibration standard to the same hydrolysis conditions as the test sample. The amino acid in free form may not completely represent the rate of destruction of labile amino acids within a peptide or protein during the hydrolysis. This is especially true for peptide bonds that are slow to cleave (e.g., Ile-Val bonds). However, this technique will allow the analyst to account for some residue destruction. Microwave acid hydrolysis has been used and is rapid but it requires special equipment as well as special precautions. The optimal conditions for microwave hydrolysis must be investigated for each individual protein/peptide sample. The microwave hydrolysis technique typically requires only a few minutes, but even a deviation of 1 minute may give inadequate results (e.g., incomplete hydrolysis or destruction of labile amino acids). Complete proteolysis, using a mixture of proteases, has been used but can be complicated, requires the proper controls, and is typically more applicable to peptides than proteins. [NOTE—During initial analyses of an unknown protein, experiments with various hydrolysis time and temperature conditions are conducted to determine the optimal conditions.]
METHOD 1
Acid hydrolysis using hydrochloric acid containing phenol is the most common procedure used for protein/peptide hydrolysis preceding amino acid analysis. The addition of phenol to the reaction prevents the halogenation of tyrosine.
Hydrolysis Solution: 6 N hydrochloric acid containing 0.1% to 1.0% of phenol.
Procedure—
Liquid Phase Hydrolysis— Place the protein or peptide sample in a hydrolysis tube, and dry. [NOTE—The sample is dried so that water in the sample will not dilute the acid used for the hydrolysis.] Add 200 µL of Hydrolysis Solution per 500 µg of lyophilized protein. Freeze the sample tube in a dry ice-acetone bath, and flame seal in vacuum. Samples are typically hydrolyzed at 110 for 24 hours in vacuum or inert atmosphere to prevent oxidation. Longer hydrolysis times (e.g., 48 and 72 hours) are investigated if there is a concern that the protein is not completely hydrolyzed.
Vapor Phase Hydrolysis— This is one of the most common acid hydrolysis procedures, and it is preferred for microanalysis when only small amounts of the sample are available. Contamination of the sample from the acid reagent is also minimized by using vapor phase hydrolysis. Place vials containing the dried samples in a vessel that contains an appropriate amount of Hydrolysis Solution. The Hydrolysis Solution does not come in contact with the test sample. Apply an inert atmosphere or vacuum (less than 200 µm of mercury or 26.7 Pa) to the headspace of the vessel, and heat to about 110 for a 24-hour hydrolysis time. Acid vapor hydrolyzes the dried sample. Any condensation of the acid in the sample vials is minimized. After hydrolysis, dry the test sample in vacuum to remove any residual acid.
METHOD 2
Tryptophan oxidation during hydrolysis is decreased by using mercaptoethanesulfonic acid (MESA) as the reducing acid.
Hydrolysis Solution: 2.5 M MESA solution.
Vapor Phase Hydrolysis— About 1 to 100 µg of the protein/peptide under test is dried in a hydrolysis tube. The hydrolysis tube is placed in a larger tube with about 200 µL of the Hydrolysis Solution. The larger tube is sealed in vacuum (about 50 µm of mercury or 6.7 Pa) to vaporize the Hydrolysis Solution. The hydrolysis tube is heated to between 170 to 185 for about 12.5 minutes. After hydrolysis, the hydrolysis tube is dried in vacuum for 15 minutes to remove the residual acid.
METHOD 3
Tryptophan oxidation during hydrolysis is prevented by using thioglycolic acid (TGA) as the reducing acid.
Hydrolysis Solution: a solution containing 7 M hydrochloric acid, 10% of trifluoroacetic acid, 20% of thioglycolic acid, and 1% of phenol.
Vapor Phase Hydrolysis— About 10 to 50 µg of the protein/peptide under test is dried in a sample tube. The sample tube is placed in a larger tube with about 200 µL of the Hydrolysis Solution. The larger tube is sealed in vacuum (about 50 µm of mercury or 6.7 Pa) to vaporize the TGA. The sample tube is heated to 166 for about 15 to 30 minutes. After hydrolysis, the sample tube is dried in vacuum for 5 minutes to remove the residual acid. Recovery of tryptophan by this method may be dependent on the amount of sample present.
METHOD 4
Cysteine-cystine and methionine oxidation is performed with performic acid before the protein hydrolysis.
Oxidation Solution— The performic acid is prepared fresh by mixing formic acid and 30 percent hydrogen peroxide (9:1), and incubated at room temperature for 1 hour.
Procedure— The protein/peptide sample is dissolved in 20 µL of formic acid, and heated at 50 for 5 minutes; then 100 µL of the Oxidation Solution is added. In this reaction, cysteine is converted to cysteic acid and methionine is converted to methionine sulfone. The oxidation is allowed to proceed for 10 to 30 minutes. The excess reagent is removed from the sample in a vacuum centrifuge. This technique may cause modifications to tyrosine residues in the presence of halides. The oxidized protein can then be acid hydrolyzed using Method 1 or Method 2.
METHOD 5
Cysteine-cystine oxidation is accomplished during the liquid phase hydrolysis with sodium azide.
Hydrolysis Solution: 6 N hydrochloric acid containing 0.2% of phenol, to which sodium azide is added to obtain a final concentration of 0.2% (w/v). The added phenol prevents halogenation of tyrosine.
Liquid Phase Hydrolysis— The protein/peptide hydrolysis is conducted at about 110 for 24 hours. During the hydrolysis, the cysteine-cystine present in the sample is converted to cysteic acid by the sodium azide present in the Hydrolysis Solution. This technique allows better tyrosine recovery than Method 4, but it is not quantitative for methionine. Methionine is converted to a mixture of the parent methionine and its two oxidative products, methionine sulfoxide and methionine sulfone.
METHOD 6
Cysteine-cystine oxidation is accomplished with dimethyl sulfoxide (DMSO).
Hydrolysis Solution: 6 N hydrochloric acid containing 0.1% to 1.0% of phenol, to which DMSO is added to obtain a final concentration of 2% (v/v).
Vapor Phase Hydrolysis— The protein/peptide hydrolysis is conducted at about 110 for 24 hours. During the hydrolysis, the cysteine-cystine present in the sample is converted to cysteic acid by the DMSO present in the Hydrolysis Solution. As an approach to limit variability and compensate for partial destruction, it is recommended to evaluate the cysteic acid recovery from oxidative hydrolyses of standard proteins containing 1 to 8 mol of cysteine. The response factors from protein/peptide hydrolysates are typically about 30% lower than those for nonhydrolyzed cysteic acid standards. Because histidine, methionine, tyrosine, and tryptophan are also modified, a complete compositional analysis is not obtained with this technique.
METHOD 7
Cysteine-cystine reduction and alkylation is accomplished by a vapor phase pyridylethylation reaction.
Reducing Solution— Transfer 83.3 µL of pyridine, 16.7 µL of 4-vinylpyridine, 16.7 µL of tributylphosphine, and 83.3 µL of water to a suitable container, and mix.
Procedure— Add the protein/peptide (between 1 and 100 µg) to a hydrolysis tube, and place in a larger tube. Transfer the Reducing Solution to the large tube, seal in vacuum (about 50 µm of mercury or 6.7 Pa), and incubate at about 100 for 5 minutes. Then remove the inner hydrolysis tube, and dry it in a vacuum desiccator for 15 minutes to remove residual reagents. The pyridylethylated protein/peptide can then be acid hydrolyzed using previously described procedures. The pyridylethylation reaction is performed simultaneously with a protein standard sample containing 1 to 8 mol of cysteine to improve accuracy in the pyridylethyl-cysteine recovery. Longer incubation times for the pyridylethylation reaction can cause modifications to the -amino terminal group and the -amino group of lysine in the protein.
METHOD 8
Cysteine-cystine reduction and alkylation is accomplished by a liquid phase pyridylethylation reaction.
Stock Solutions— Prepare and filter three solutions: 1 M Tris hydrochloride (pH 8.5) containing 4 mM edetate disodium (Stock Solution 1), 8 M guanidine hydrochloride (Stock Solution 2), and 10% of 2-mercaptoethanol in water (Stock Solution 3).
Reducing Solution— Prepare a mixture of Stock Solution 2 and Stock Solution 1 (3:1) to obtain a buffered solution of 6 M guanidine hydrochloride in 0.25 M Tris hydrochloride.
Procedure— Dissolve about 10 µg of the test sample in 50 µL of the Reducing Solution, and add about 2.5 µL of Stock Solution 3. Store under nitrogen or argon for 2 hours at room temperature in the dark. To achieve the pyridylethylation reaction, add about 2 µL of 4-vinylpyridine to the protein solution, and incubate for an additional 2 hours at room temperature in the dark. The protein/peptide is desalted by collecting the protein/peptide fraction from a reverse-phase HPLC separation. The collected sample can be dried in a vacuum centrifuge before acid hydrolysis.
METHOD 9
Cysteine-cystine reduction and alkylation is accomplished by a liquid phase carboxymethylation reaction.
Stock Solutions— Prepare as directed for Method 8.
Carboxymethylation Solution— Prepare a solution containing 100 mg of iodoacetamide per mL of alcohol.
Buffer Solution— Use the Reducing Solution, prepared as directed for Method 8.
Procedure— Dissolve the test sample in 50 µL of the Buffer Solution, and add about 2.5 µL of Stock Solution 3. Store under nitrogen or argon for 2 hours at room temperature in the dark. Add the Carboxymethylation Solution in a 1.5 fold ratio per total theoretical content of thiols, and incubate for an additional 30 minutes at room temperature in the dark. [NOTE—If the thiol content of the protein is unknown, then add 5 µL of 100 mM iodoacetamide for every 20 nmol of protein present.] The reaction is stopped by adding excess of 2-mercaptoethanol. The protein/peptide is desalted by collecting the protein/peptide fraction from a reverse-phase HPLC separation. The collected sample can be dried in a vacuum centrifuge before acid hydrolysis. The S-carboxyamidomethylcysteine formed will be converted to S-carboxymethyl-cysteine during acid hydrolysis.
METHOD 10
Cysteine-cystine is reacted with dithiodiglycolic acid or dithiodipropionic acid to produce a mixed disulfide. [NOTE—The choice of dithiodiglycolic acid or dithiodipropionic acid depends on the required resolution of the amino acid analysis method.]
Reducing Solution: a solution containing 10 mg of dithiodiglycolic acid (or dithiodipropionic acid) per mL of 0.2 M sodium hydroxide.
Procedure— Transfer about 20 µg of the test sample to a hydrolysis tube, and add 5 µL of the Reducing Solution. Add 10 µL of isopropyl alcohol, and then remove all of the sample liquid by vacuum centrifugation. The sample is then hydrolyzed using Method 1. This method has the advantage that other amino acid residues are not derivatized by side reactions, and the sample does not need to be desalted prior to hydrolysis.
METHOD 11
Asparagine and glutamine are converted to aspartic acid and glutamic acid, respectively, during acid hydrolysis. Asparagine and aspartic acid residues are added and represented by Asx, while glutamine and glutamic acid residues are added and represented by Glx. Proteins/peptides can be reacted with bis(1,1-trifluoroacetoxy)iodobenzene (BTI) to convert the asparagine and glutamine residues to diaminopropionic acid and diaminobutyric acid residues, respectively, upon acid hydrolysis. These conversions allow the analyst to determine the asparagine and glutamine content of a protein/peptide in the presence of aspartic acid and glutamic acid residues.
Reducing Solutions— Prepare and filter three solutions: a solution of 10 mM trifluoroacetic acid (Solution 1), a solution of 5 M guanidine hydrochloride and 10 mM trifluoroacetic acid (Solution 2), and a freshly prepared solution of dimethylformamide containing 36 mg of BTI per mL (Solution 3).
Procedure— In a clean hydrolysis tube, transfer about 200 µg of the test sample, and add 2 mL of Solution 1 or Solution 2 and 2 mL of Solution 3. Seal the hydrolysis tube in vacuum. Heat the sample at 60 for 4 hours in the dark. The sample is then dialyzed with water to remove the excess reagents. Extract the dialyzed sample three times with equal volumes of n-butyl acetate, and then lyophilize. The protein can then be acid hydrolyzed using previously described procedures. The -, -diaminopropionic and -, -diaminobutyric acid residues do not typically resolve from the lysine residues upon ion-exchange chromatography based on amino acid analysis. Therefore, when using ion-exchange as the mode of amino acid separation, the asparagine and glutamine contents are the quantitative difference in the aspartic acid and glutamic acid assayed contents with un-derivatized and BTI-derivatized acid hydrolysis. [NOTE—The threonine, methionine, cysteine, tyrosine, and histidine assayed content can be altered by BTI derivatization; a hydrolysis without BTI will have to be performed if the analyst is interested in the protein/peptide content of these other amino acid residues.]
Methodologies of Amino Acid Analysis
Many amino acid analysis techniques exist, and the choice of any one technique often depends on the sensitivity required from the assay. In general, about one-half of the amino acid analysis techniques employed rely on the separation of the free amino acids by ion-exchange chromatography followed by postcolumn derivatization (e.g., with ninhydrin or o-phthalaldehyde). Postcolumn detection techniques can be used with samples that contain small amounts of buffer components, such as salts and urea, and generally require between 5 and 10 µg of protein sample per analysis. The remaining amino acid techniques typically involve precolumn derivatization of the free amino acids (e.g., phenyl isothiocyanate; 6-aminoquinolyl-N-hydroxysuccinimidyl carbonate; (dimethylamino)azobenzenesulfonyl chloride; 9-fluorenyl-methylchloroformate; and 7-fluoro-4-nitrobenzo-2-oxa-1,3-diazole) followed by reverse-phase HPLC. Precolumn derivatization techniques are very sensitive and usually require between 0.5 and 1.0 µg of protein sample per analysis but may be influenced by buffer salts in the samples. Precolumn derivatization techniques may also result in multiple derivatives of a given amino acid, which complicates the result interpretation. Postcolumn derivatization techniques are generally influenced less by performance variation of the assay than precolumn derivatization techniques.
The following Methods may be used for quantitative amino acid analysis. Instruments and reagents for these procedures are available commercially. Furthermore, many modifications of these methodologies exist with different reagent preparations, reaction procedures, chromatographic systems, etc. Specific parameters may vary according to the exact equipment and procedure used. Many laboratories will utilize more than one amino acid analysis technique to exploit the advantages offered by each. In each of these Methods, the analog signal is visualized by means of a data acquisition system, and the peak areas are integrated for quantification purposes.
METHOD 1—POSTCOLUMN NINHYDRIN DETECTION
Ion-exchange chromatography with postcolumn ninhydrin detection is one of the most common methods employed for quantitative amino acid analysis. As a rule, a Li-based cation-exchange system is employed for the analysis of the more complex physiological samples, and the faster Na-based cation-exchange system is used for the more simplistic amino acid mixtures obtained with protein hydrolysates (typically containing 17 amino acid components). Separation of the amino acids on an ion-exchange column is accomplished through a combination of changes in pH and cation strength. A temperature gradient is often employed to enhance separation.
When the amino acid reacts with ninhydrin, the reactant has characteristic purple or yellow color. Amino acids, except imino acids, give a purple color, and show maximum absorption at 570 nm. The imino acids, such as proline, give a yellow color, and show maximum absorption at 440 nm. The postcolumn reaction between ninhydrin and amino acid eluted from the column is monitored at 440 nm and 570 nm, and the chromatogram obtained is used for the determination of amino acid composition.
The detection limit is considered to be 10 pmol for most of the amino acid derivatives, but 50 pmol for proline. Response linearity is obtained in the range of 20 to 500 pmol with correlation coefficients exceeding 0.999. To obtain good compositional data, samples larger than 1 µg before hydrolysis are best suited for this amino acid analysis of protein/peptide.
One method for postcolumn ninhydrin detection is shown below. Many other methods are also available, with instruments and reagents available commercially.
Mobile Phase Preparation—
Solution A— Transfer about 1.7 g of anhydrous sodium citrate and 1.5 mL of hydrochloric acid to a 100-mL volumetric flask, dissolve in and dilute with water to volume, and mix. Adjust, if necessary, with hydrochloric acid to a pH of 3.0.
Solution B— Transfer about 1.7 g of anhydrous sodium citrate and 0.7 mL of hydrochloric acid to a 100-mL volumetric flask, dissolve in and dilute with water to volume, and mix. Adjust, if necessary, with hydrochloric acid to a pH of 4.3.
Solution C— Prepare a solution containing 5% of sodium chloride, 1.9% of anhydrous sodium citrate, and 0.1% of phenol in water, and adjust to a pH of 6.
Column Regeneration Solution— Prepare a solution containing 0.8% of sodium hydroxide in water, and adjust to a pH of 13.
Mobile Phase— Use variable mixtures of Solution A, Solution B, and Solution C as directed for Chromatographic system.
Postcolumn Reagent— Transfer about 18 g of ninhydrin and 0.7 g of hydrindantin to 900 mL of a solution containing 76.7% of dimethyl sulfoxide, 0.7% of dihydrate lithium acetate, and 0.1% of acetic acid, and mix for at least 3 hours under inert gas, such as nitrogen. [NOTE—This reagent is stable for 30 days if kept between 2 and 8 under inert gas.]
Buffer Solution— Prepare a solution containing 2% of anhydrous sodium citrate, 1% of hydrochloric acid, 0.5% of thiodiglycol, and 0.1% of benzoic acid in water, and adjust to a pH of 2.
Chromatographic System— The liquid chromatograph is equipped with a detector with appropriate interference filters at 440, 570, or 690 nm and a 4.0-mm × 120-mm column that contains 7.5-µm sulfonated styrene-divinylbenzene copolymer packing. The flow rate is about 14 mL per hour. The system is programmed as follows. Initially equilibrate the column with Solution A; at 25 minutes, the composition of the Mobile Phase is changed to 100% Solution B; and at 37 minutes, the composition is changed to 100% Solution C. At 75 minutes into the run, the last amino acid has been eluted from the column, and the column is regenerated with Column Regeneration Solution for 1 minute. The column is then equilibrated with Solution A for 11 minutes before the next injection. The column temperature is programmed as follows. The initial temperature is 48; after 11.5 minutes, the temperature is increased to 65 at a rate of 3 per minute; at about 35 minutes, the temperature is increased to 77 at a rate of 3 per minute; and finally at about 52 minutes, the temperature is decreased to 48 at a rate of 3 per minute.
Procedure and Postcolumn Reaction— Reconstitute the lyophilized protein/peptide hydrolysate in the Buffer Solution, inject an appropriate amount into the chromatograph, and proceed as directed for Chromatographic System. As the amino acids are eluted from the column, they are mixed with the Postcolumn Reagent, which is delivered at a flow rate of 7 mL per hour, through a tee. After mixing, the column effluent and the Postcolumn Reagent pass through a tubular reactor at a temperature of 135, where a characteristic purple or yellow color is developed. From the reactor, the liquid passes through a colorimeter with a 12-mm flow-through cuvette. The light emerging from the cuvette is split into three beams for analysis by the detector with interference filters at 440, 570, or 690 nm. The 690-nm signal may be electronically subtracted from the other signals for improved signal-to-noise ratios. The 440-nm (imino acids) and the 570-nm (amino acids) signals may be added in order to simplify data handling.
METHOD 2—POSTCOLUMN OPA FLUOROMETRIC DERIVATIZATION
Ion-exchange chromatography with postcolumn o-phthalaldehyde (OPA) fluorometric detection is used. The procedure employs an ion-exchange column for separation of free amino acids followed by postcolumn oxidation with sodium hypochlorite and derivatization using OPA and N-acetyl-L-cysteine. The sodium hypochlorite oxidation step allows secondary amines, such as proline, to react with the OPA reagent.
OPA reacts with primary amines in the presence of thiol compound to form highly fluorescent isoindole products. This reaction is utilized for the postcolumn derivatization in analysis of amino acids by ion-exchange chromatography. The rule of the separation is the same as Method I. Instruments and reagents for this form of amino acid analysis are available commercially. Many modifications of this method exist.
Although OPA does not react with secondary amines (imino acids, such as proline) to form fluorescent substances, the oxidation with sodium hypochlorite allows secondary amines to react with OPA. The procedure employs a strongly acidic cation-exchange column for separation of free amino acids followed by postcolumn oxidation with sodium hypochlorite and postcolumn derivatization using OPA and thiol compound, such as N-acetyl-L-cysteine and 2-mercaptoethanol. The derivatization of primary amino acids are not noticeably affected by the continuous supply of sodium hypochlorite.
Separation of the amino acids on an ion-exchange column is accomplished through a combination of changes of pH and cation strength. After postcolumn derivatization of eluted amino acids with OPA, the reactant passes through the fluorometric detector. Fluorescence intensity of OPA-derivatized amino acids are monitored with an excitation wavelength of 348 nm and an emission wavelength of 450 nm.
The detection limit is considered to be a few tens of picomole level for most of the amino acid derivatives. Response linearity is obtained in the range of a few picomole level to a few tens of nanomole level. To obtain good compositional data, a sample greater than 500 ng before hydrolysis is best suited for the amino acid analysis of protein/peptide.
One method of postcolumn OPA fluorometric detection is shown below.
Mobile Phase Preparation—
Solution A— Prepare a solution of sodium hydroxide, citric acid, and alcohol in HPLC grade water having a 0.2 N sodium concentration and containing 7% of alcohol (w/v), adjusted to a pH of 3.2.
Solution B— Prepare a solution of sodium hydroxide and citric acid in HPLC grade water having a 0.6 N sodium concentration, adjusted to a pH of 10.0.
Solution C: 0.2 N sodium hydroxide.
Mobile Phase— Use variable mixtures of Solution A, Solution B, and Solution C as directed for Chromatographic System.
Postcolumn Reagent Preparation—
Alkaline Buffer— Prepare a solution containing 384 mM sodium carbonate, 216 mM boric acid, and 108 mM potassium sulfate, and adjust to a pH of 10.0.
Hypochlorite Reagent— To 1 L of Alkaline Buffer, add 0.4 mL of sodium hypochlorite solution (10% chlorine concentration). [NOTE—The hypochlorite solution is stable for 2 weeks.]
OPA Reagent— Transfer 2 g of N-acetyl-L-cysteine and 1.6 g of OPA to a 15-mL volumetric flask, dissolve in and dilute with alcohol to volume, and mix. Transfer this solution and 4 mL of 10% aqueous polyethylene (23) lauryl ether2 to a 1-liter volumetric flask, dilute with 980 mL of Alkaline Buffer, and mix.
Chromatographic System— The liquid chromatograph is equipped with a fluorometric detector set to an excitation wavelength of 348 nm and an emission wavelength of 450 nm and a 4.0-mm × 150-mm column that contains 7.5-µm packing L17. The flow rate is about 0.3 mL per minute, and the column temperature is set at 50. The system is programmed as follows. The column is equilibrated with Solution A; over the next 20 minutes, the composition of the Mobile Phase is changed linearly to 85% Solution A and 15% Solution B; then there is a step change to 40% Solution A and 60% Solution B; over the next 18 minutes, the composition is changed linearly to 100% Solution B and held for 7 minutes; then there is a step change to 100% Solution C, and this is held for 6 minutes; then there is a step change to Solution A, and this composition is maintained for the next 8 minutes.
Procedure and Postcolumn Reaction— Inject about 1.0 nmol of each amino acid under test into the chromatograph, and proceed as directed for Chromatographic System. As the effluent leaves the column, it is mixed with the Hypochlorite Reagent. The mixture passes through the first postcolumn reactor which consists of stainless steel 0.5-mm × 2-m tubing. A second postcolumn reactor of similar design is placed immediately downstream from the first postcolumn reactor and is used for the OPA postcolumn reaction. The flow rates for both the Hypochlorite Reagent and the OPA Reagent are 0.2 mL per minute, resulting in a total flow rate (i.e., Hypochlorite Reagent, OPA Reagent, and column effluent) of 0.7 mL per minute exiting from the postcolumn reactors. Postcolumn reactions are conducted at 55. This results in a residence time of about 33 seconds in the OPA postcolumn reactor. After postcolumn derivatization, the column effluent passes through the fluorometric detector.
METHOD 3—PRECOLUMN DETERMINATION
Precolumn derivatization of amino acids with phenylisothiocyanate (PITC) followed by reverse-phase HPLC separation with UV detection is used.
PITC reacts with amino acids to form phenylthiocarbamyl (PTC) derivatives which can be detected with high sensitivity at 254 nm. Therefore, precolumn derivatization of amino acids with PITC followed by a reverse-phase HPLC separation with UV detection is used to analyze the amino acid composition.
After the reagent is removed under vacuum, the derivatized amino acids can be stored dry and frozen for several weeks with no significant degradation. If the solution for injection is kept cold, no noticeable loss in chromatographic response occurs after three days.
Separation of the PTC-amino acids on a reverse-phase HPLC with ODS column is accomplished through a combination of changes in concentrations of acetonitrile and buffer ionic strength. PTC-amino acids eluted from the column are monitored at 254 nm.
The detection limit is considered to be 1 pmol for most of the amino acid derivatives. Response linearity is obtained in the range of 20 to 500 pmol with correlation coefficients exceeding 0.999. To obtain good compositional data, a sample larger than 500 ng of protein/peptide before hydrolysis is best suited for this amino analysis of proteins/peptides.
One method of precolumn PITC derivatization is described below.
Mobile Phase Preparation—
Solution A: 0.05 M ammonium acetate, adjusted with phosphoric acid to a pH of 6.8.
Solution B— Prepare 0.1 M ammonium acetate, adjust with phosphoric acid to a pH of 6.8, and then prepare a mixture of this solution and acetonitrile (1:1).
Solution C: a mixture of acetonitrile and water (70:30).
Mobile Phase— Use variable mixtures of Solution A, Solution B, and Solution C as directed for Chromatographic System.
Derivatization Reagent Preparation—
Coupling Buffer: a mixture of acetonitrile, pyridine, triethylamine, and water (10:5:2:3).
Sample Solvent: a mixture of water and acetonitrile (7:2).
Sample Derivatization Procedure— Dissolve the lyophilized test sample in 100 µL of the Coupling Buffer, and then dry in a vacuum centrifuge to remove any hydrochloride if a protein hydrolysis step was used. Dissolve the test sample in 100 µL of Coupling Buffer, add 5 µL of PITC, and incubate at room temperature for 5 minutes. The test sample is again dried in a vacuum centrifuge, and is dissolved in 250 µL of Sample Solvent.
Chromatographic System— The liquid chromatograph is equipped with a 254-nm detector and a 4.6-mm × 250-mm column that contains 5-µm packing L1. The flow rate is about 1 mL per minute, and the column temperature is maintained at 52. The system is programmed as follows. The column is equilibrated with Solution A; over the next 15 minutes, the composition of the Mobile Phase is changed linearly to 85% Solution A and 15% Solution B; over the next 15 minutes, the composition is changed linearly to 50% Solution A and 50% Solution B; then there is a step change to 100% Solution C, and this is held for 10 minutes; then there is a step change to 100% Solution A, and the column is allowed to equilibrate before the next injection.
Procedure— Inject about 1.0 nmol of each PITC-amino acid under test (10-µL sample in Sample Solvent) into the chromatograph, and proceed as directed for Chromatographic System.
METHOD 4—PRECOLUMN AQC DERIVATIZATION
Precolumn derivatization of amino acids with 6-aminoquinolyl-N-hydroxysuccinimidyl carbamate (AQC) followed by reverse-phase HPLC separation with fluorometric detection is used.
AQC reacts with amino acids to form stable, fluorescent unsymmetric urea derivatives (AQC-amino acids) which are readily amenable to analysis by reverse-phase HPLC. Therefore, precolumn derivatization of amino acids with AQC followed by reverse-phase HPLC separation is used to analyze the amino acid composition.
Separation of the AQC-amino acids on an ODS column is accomplished through a combination of changes in the concentrations of acetonitrile and salt. Selective fluorescence detection of the derivatives with an excitation wavelength at 250 nm and an emission wavelength at 395 nm allows for the direct injection of the reaction mixture with no significant interference from the only major fluorescent reagent by-product, 6-aminoquinoline. Excess reagent is rapidly hydrolyzed (t1/2< 15 seconds) to yield 6-aminoquinoline-N-hydroxysuccinimide and carbon dioxide, and after 1 minute no further derivatization can take place.
Peak areas for AQC-amino acids are essentially unchanged for at least 1 week at room temperature, and the derivatives have more than sufficient stability to allow for overnight automated chromatographic analysis.
The detection limit is considered to be ranging from about 40 fmol to 320 fmol for each amino acid, except for Cys. The detection limit for Cys is approximately 800 fmol. Response linearity is obtained in the range of 2.5 µM to 200 µM with correlation coefficients exceeding 0.999. Good compositional data can be obtained from the analysis of derivatized protein hydrolysates containing as little as 30 ng of protein/peptide.
One method of precolumn AQC derivatization is shown below.
Mobile Phase Preparation—
Solution A— Prepare a solution having a composition of 140 mM sodium acetate and 17 mM triethylamine, and adjust with phosphoric acid to a pH of 5.02.
Solution B: a mixture of acetonitrile and water (60:40).
Mobile Phase— Use variable mixtures of Solution A and Solution B as directed for Chromatographic System.
Sample Derivatization Procedure— Dissolve about 2 µg of the test sample in 20 µL of 15 mM hydrochloric acid, and dilute with 0.2 M borate buffer (pH 8.8) to 80 µL. The derivatization is initiated by the addition of 20 µL of 10 mM AQC in acetonitrile, and allowed to proceed for 10 minutes at room temperature.
Chromatographic System— The liquid chromatograph is equipped with a fluorometric detector set at an excitation wavelength of 250 nm and an emission wavelength of 395 nm and a 3.9-mm × 150-mm column that contains 4-µm packing L1. The flow rate is about 1 mL per minute, and the column temperature is maintained at 37. The system is programmed as follows. The column is equilibrated with Solution A; over the next 0.5 minute, the composition of the Mobile Phase is changed linearly to 98% Solution A and 2% Solution B; then over the next 14.5 minutes to 93% Solution A and 7% Solution B; then over the next 4 minutes to 87% Solution A and 13% Solution B; over the next 14 minutes to 68% Solution A and 32% Solution B; then there is a step change to 100% Solution B for a 5-minute wash; over the next 10 minutes, there is a step change to 100% Solution A; and the column is allowed to equilibrate before the next injection.
Procedure— Inject about 0.05 nmol of each AQC-amino acid under test into the chromatograph, and proceed as directed for Chromatographic System.
METHOD 5—PRECOLUMN OPA DERIVATIZATION
Precolumn derivatization of amino acids with OPA followed by reverse-phase HPLC separation with fluorometric detection is used. This technique does not detect amino acids that exist as secondary amines (e.g., proline).
OPA in conjunction with a thiol reagent reacts with primary amine groups to form highly fluorescent isoindole products. 2-Mercaptoethanol and 3-mercaptopropionic acid can be used as thiol. OPA itself does not fluoresce and consequently produces no interfering peaks. In addition, its solubility and stability in aqueous solution, along with the rapid kinetics for the reactions, make it amenable to automated derivatization and analysis using an autosampler to mix the sample with the reagent. However, lack of reactivity with secondary amino acids has been a predominant drawback. This method does not detect amino acids that exist as secondary amines (e.g., proline). To compensate for this drawback, this technique may be combined with another technique described in Method 7 or Method 8.
Precolumn derivatization of amino acids with OPA is followed by reverse-phase HPLC separation. Because of the instability of the OPA-amino acid derivative, HPLC separation and analysis are performed immediately following derivatization. The liquid chromatograph is equipped with a fluorometric detector for the detection of derivatized amino acids. Fluorescence intensity of the OPA-derivatized amino acids are monitored with an excitation wavelength of 348 nm and an emission wavelength of 450 nm.
The detection limits as low as 50 fmol via fluorescence have been reported, although the practical limit of analysis remains at 1 pmol. One method of precolumn OPA derivatization is shown below.
Mobile Phase Preparation—
Solution A: a mixture of 100 mM sodium acetate (pH 7.2), methanol, and tetrahydrofuran (900:95:5).
Solution B: methanol.
Mobile Phase— Use variable mixtures of Solution A and Solution B as directed for Chromatographic System.
Derivatization Reagent— Dissolve 50 mg of OPA in 1.25 mL of methanol (protein sequencing grade). Add 50 µL of 2-mercaptoethanol and 11.2 mL of 0.4 M sodium borate (pH 9.5), and mix. [NOTE—This reagent is stable for 1 week.]
Sample Derivatization Procedure— Transfer about 5 µL of the test sample to an appropriate container, add 5 µL of the Derivatization Reagent, and mix. After 1 minute, add not less than 20 µL of 0.1 M sodium acetate (pH 7.0). Use 20 µL of this solution for analysis. [NOTE—Use of an internal standard (e.g., norleucine) is recommended for quantitative analysis because of potential reagent volume variations in the sample derivatization. The sample derivatization is performed in an automated on-line fashion. Because of the instability of the OPA-amino acid derivative, HPLC separation and analysis are performed immediately following derivatization.]
Chromatographic System— The liquid chromatograph is equipped with a fluorometric detector set at an excitation wavelength of 348 nm and an emission wavelength of 450 nm and a 4.6-mm × 75-mm column that contains 3-µm packing L3. The flow rate is about 1.7 mL per minute, and the column temperature is maintained at 37. The system is programmed as follows. The column is equilibrated with 92% Solution A and 8% Solution B; over the next 2 minutes, the composition of the Mobile Phase is changed to 83% Solution A and 17% Solution B, and held for an additional 3 minutes; then changed to 54% Solution A and 46% Solution B over the next 5 minutes, and held for an additional 2 minutes; then changed to 34% Solution A and 66% Solution B over the next 2 minutes, and held for 1 minute; then over the next 0.3 minute changed to 20% Solution A and 80% Solution B, and held for an additional 2.6 minutes; and then finally over 0.6 minute changed to 92% Solution A and 8% Solution B, and held for an additional 0.6 minute.
Procedure— Inject about 0.02 nmol of each OPA-amino acid under test into the chromatograph, and proceed as directed for Chromatographic System.
METHOD 6—POSTCOLUMN DABS-Cl DERIVATIZATION
Precolumn derivatization of amino acids with (dimethylamino)azobenzenesulfonyl chloride (DABS-Cl) followed by reverse-phase HPLC separation with visible light detection is used.
DABS-Cl is a chromophoric reagent employed for the labeling of amino acids. Amino acids labeled with DABS-Cl (DABS-amino acids) are highly stable and show the maximum absorption at 436 nm.
DABS-amino acids, all 19 naturally occurring amino acids derivatives, can be separated on an ODS column of a reverse-phase HPLC by employing gradient systems consisting of acetonitrile and aqueous buffer mixture. Separated DABS-amino acids eluted from the column are detected at 436 nm in the visible region.
This method can analyze the imino acids, such as proline, together with the amino acids, at the same degree of sensitivity. DABS-Cl derivatization method permits the simultaneous quantification of tryptophan residues by previous hydrolysis of the protein/peptide with sulfonic acids, such as mercaptoethanesulfonic acid, p-toluenesulfonic acid, or methanesulfonic acid, described for Method 2 in Protein Hydrolysis under Amino Acid Analysis. The other acid-labile residues, asparagine and glutamine, can also be analyzed by previous conversion into diaminopropionic acid and diaminobutyric acid, respectively, by treatment of protein/peptide with BTI, described for Method 11 in Protein Hydrolysis under Amino Acid Analysis.
The non-proteinogenic amino acid, norleucine, cannot be used as an internal standard in this method as this compound is eluted in a chromatographic region crowded with peaks of primary amino acids. Nitrotyrosine can be used as an internal standard because it is eluted in a clean region.
The detection limit of DABS-amino acid is about 1 pmol. As little as 2 to 5 pmol of an individual DABS-amino acid can be quantitatively analyzed with reliability, and only 10 ng to 30 ng of the dabsylated protein hydrolysate is required for each analysis.
One method for precolumn DABS-Cl derivatization is shown below.
Mobile Phase Preparation—
Solution A: 25 mM sodium acetate (pH 6.5) containing 4% of dimethylformamide.
Solution B: acetonitrile.
Mobile Phase— Use variable mixtures of Solution A and Solution B as directed for Chromatographic System.
Derivatization Reagent Preparation—
Sample Buffer: 50 mM sodium bicarbonate, adjusted to a pH of 8.1.
Derivatization Reagent— Dissolve 1.3 mg of DABS-Cl in 1 mL of acetonitrile. [NOTE—This reagent is prepared fresh shortly before the derivatization step.]
Sample Dilution Buffer— Prepare a mixture of 50 mM sodium phosphate (pH 7.0) and alcohol (1:1).
Sample Derivatization Procedure— Dissolve the test sample in 20 µL of Sample Buffer, add 40 µL of Derivatization Reagent, and mix. The sample container is sealed with a silicon-rubber stopper, and heated to 70 for 10 minutes. During the sample heating, the mixture will become completely soluble. After the derivatization, dilute the test sample with an appropriate quantity of the Sample Dilution Buffer.
Chromatographic System— The liquid chromatograph is equipped with a 436-nm detector and a 4.6-mm × 250-mm column that contains packing L1. The flow rate is about 1 mL per minute, and the column temperature is maintained at 40. The system is programmed as follows. The column is equilibrated with 85% Solution A and 15% Solution B; over the next 20 minutes, the composition of the Mobile Phase is changed to 60% Solution A and 40% Solution B; over the next 12 minutes, the composition is changed to 30% Solution A and 70% Solution B, and held for an additional 2 minutes.
Procedure— Inject about 0.05 nmol of the DABS-amino acids into the chromatograph, and proceed as directed for Chromatographic System.
METHOD 7—PRECOLUMN FMOC-Cl DERIVATIZATION
Precolumn derivatization of amino acids with 9-fluorenylmethyl chloroformate (FMOC-Cl) followed by reverse-phase HPLC separation with fluorometric detection is used.
FMOC-Cl reacts with both primary and secondary amino acids to form highly fluorescent products. The reaction of FMOC-Cl with amino acid proceeds under mild conditions, in aqueous solution, and is completed in 30 seconds. The derivatives are stable, with only the histidine derivative showing any breakdown. Although FMOC-Cl is fluorescent itself, the reagent excess and fluorescent side-products can be eliminated without loss of FMOC-amino acids.
FMOC-amino acids are separated by reverse-phase HPLC using an ODS column. The separation is carried out by gradient elution varied linearly from a mixture of acetic acid buffer, methanol, and acetonitrile (50:40:10) to a mixture of acetonitrile and acetic acid buffer (50:50), and 20 amino acid derivatives that are separated in 20 minutes. Each derivative eluted from the column is monitored by a fluorometric detector set at an excitation wavelength of 260 nm and an emission wavelength of 313 nm.
The detection limit is in the low fmol range. A linearity range of 0.1 µM to 50 µM is obtained for most amino acids.
One method for precolumn FMOC-Cl derivatization is shown below.
Mobile Phase Preparation—
Acetic Acid Buffer— Transfer 3 mL of glacial acetic acid and 1 mL of triethylamine to a 1-liter volumetric flask, and dilute with HPLC grade water to volume. Adjust with sodium hydroxide to a pH of 4.20.
Solution A: a mixture of Acetic Acid Buffer, methanol, and acetonitrile (50:40:10).
Solution B: a mixture of acetonitrile and Acetic Acid Buffer (50:50).
Mobile Phase— Use variable mixtures of Solution A and Solution B as directed for Chromatographic System.
Derivatization Reagent Preparation
Borate Buffer— Prepare a 1 M boric acid solution, and adjust with sodium hydroxide to a pH of 6.2.
FMOC-Cl Reagent— Dissolve 155 mg of 9-fluorenylmethyl chloroformate in 40 mL of acetone, and mix.
Sample Derivatization Procedure— To 0.4 mL of the test sample add 0.1 mL of Borate Buffer and 0.5 mL of FMOC-Cl Reagent. After about 40 seconds, extract the mixture with 2 mL of pentane, and then extract again with fresh pentane. The aqueous solution with amino acid derivatives is then ready for injection.
Chromatographic System— The liquid chromatograph is equipped with a fluorometric detector set at an excitation wavelength of 260 nm and an emission wavelength of 313 nm and a 4.6-mm × 125-mm column that contains 3-µm packing L1. The flow rate is about 1.3 mL per minute. The system is programmed as follows. The column is equilibrated with Solution A, and this composition is maintained for 3 minutes; over the next 9 minutes, it is changed to 100% Solution B; then over the next 0.5 minute, the flow rate is increased to 2 mL per minute, and held until the final FMOC-amino acid is eluted from the column. The total run time is about 20 minutes.
Procedure— Inject not less than 0.01 nmol of each FMOC-amino acid under test into the chromatograph, and proceed as directed for Chromatographic System. The FMOC-histidine derivative will generally give a lower response than the other derivatives.
METHOD 8—PRECOLUMN NBD-F DERIVATIZATION
Precolumn derivatization of amino acids with 7-fluoro-4-nitrobenzo-2-oxa-1,3-diazole (NBD-F) followed by reverse-phase HPLC separation with fluorometric detection is used.
7-Fluoro-4-nitrobenzo-2-oxa-1,3-diazole (NBD-F) reacts with both primary and secondary amino acids to form highly fluorescent products. Amino acids are derivatized with NBD-F by heating to 60 for 5 minutes.
NBD-amino acid derivatives are separated on an ODS column of reverse-phase HPLC by employing a gradient elution system consisting of acetonitrile and aqueous buffer mixture, and 17 amino acid derivatives that are separated in 35 minutes. E-aminocaproic acid can be used as an internal standard because it is eluted in a clean chromatographic region. Each derivative eluted from the column is monitored by a fluorometric detector set at an excitation wavelength of 480 nm and an emission wavelength of 530 nm.
The sensitivity of this method is almost the same as that for the precolumn OPA derivatization method (Method 5), excluding proline to which OPA is not reactive and might be advantageous for NBD-F against OPA.
The detection limit for each amino acid is about 10 fmol. Profile analysis was achieved for about 1.5 mg of protein hydrolysates in the final precolumn labeling reaction mixture for HPLC.
One method for precolumn NBD-F derivatization is shown below.
Mobile Phase Preparation—
Solution A: a solution of 10mM sodium citrate containing 75 mM sodium perchlorate, adjusted with hydrochloric acid to a pH of 6.2.
Solution B: a mixture of acetonitrile and water (50:50).
Derivatization Reagent Preparation—
Sample Buffer: a 0.1 M boric acid solution, adjusted with sodium hydroxide to a pH of 9.2.
Derivatization Reagent— Dissolve 5 mg of NBD-F in 1.0 mL of alcohol, and mix.
Sample Derivatization Procedure— Dissolve the test sample in 20 µL of Sample buffer, add 10 µL of Derivatization Reagent, and mix. The sample container is heated at 60 for 5 minutes. After the derivatization, dilute the test sample with 300 µL of Solution A.
Chromatographic System— The liquid chromatograph is equipped with a fluorometric detector set at an excitation wavelength of 480 nm and an emission wavelength of 530 nm and a 4.6-mm × 150-mm column that contains 5-µm particle size ODS silica packing. The flow rate is about 1.0 mL per minute, and the column temperature is maintained at 40. The system is programmed as follows. The column is equilibrated with 94% Solution A and 6% Solution B; over the next 16 minutes, the composition is changed linearly to 63% Solution A and 37% Solution B; over the next 5 minutes, the composition is changed linearly to 62% Solution A and 38% Solution B; over the next 9 minutes, the composition is changed linearly to 100% Solution B, and held for an additional 5 minutes; then finally over 2 minutes, the composition is changed linearly to 94% Solution A and 6% Solution B; and then the column is allowed to equilibrate before the next injection.
Procedure— Inject about 15 pmol of each NBD-amino acid under test into the chromatograph, and proceed as directed for Chromatographic System.
Data Calculation and Analysis
When determining the amino acid content of a protein/peptide hydrolysate, it should be noted that the acid hydrolysis step destroys tryptophan and cysteine. Serine and threonine are partially destroyed by acid hydrolysis, while isoleucine and valine residues may be only partially cleaved. Methionine can undergo oxidation during acid hydrolysis, and some amino acids (e.g., glycine and serine) are common contaminants. Application of adequate vacuum (less than 200 um µm of mercury or 26.7 Pa) or introduction of inert gas (argon) in the headspace of the reaction vessel during vapor phase hydrolysis can reduce the level of oxidative destruction. Therefore, the quantitative results obtained for cysteine, tryptophan, threonine, isoleucine, valine, methionine, glycine, and serine from a protein/peptide hydrolysate may be variable and may warrant further investigation and consideration.
CALCULATIONS
Amino Acid Mole Percent— This is the number of specific amino acid residues per 100 residues in a protein. This result may be useful for evaluating amino acid analysis data when the molecular weight of the protein/peptide under investigation is unknown. This information can be used to corroborate the identity of a protein and has other applications. Carefully identify and integrate the peaks obtained as directed for each Procedure. Calculate the mole percent for each amino acid present in the test sample by the formula:
100rU / r,
in which rU is the peak response, in nmol, of the amino acid under test; and r is the sum of peak responses, in nmol, for all amino acids present in the test sample. Comparison of the mole percent of the amino acids under test to data from known proteins can help establish or corroborate the identity of the sample protein.
Unknown Protein Samples— This data analysis technique can be used to estimate the protein concentration of an unknown protein sample using the amino acid analysis data. Calculate the mass, in µg, of each recovered amino acid by the formula:
mMW/1000,
in which m is the recovered quantity, in nmol, of the amino acid under test; and MW is the average molecular weight, in mg, for that amino acid, corrected for the weight of the water molecule that was eliminated during peptide bond formation. The sum of the masses of the recovered amino acids will give an estimate of the total mass of the protein analyzed after appropriate correction for partially and completely destroyed amino acids. If the molecular weight of the unknown protein is available (i.e., by SDS-PAGE analysis or mass spectroscopy), the amino acid composition of the unknown protein can be predicted. Calculate the number of residues of each amino acid by the formula:
m/(1000M/MWT),
in which m is the recovered quantity, in nmol, of the amino acid under test; M is the total mass, in µg, of the protein; and MWT is the molecular weight, in mg, of the unknown protein.
Known Protein Samples— This data analysis technique can be used to investigate the amino acid composition and protein concentration of a protein sample of known molecular weight and amino acid composition using the amino acid analysis data. When the composition of the protein being analyzed is known, one can exploit the fact that some amino acids are recovered well, while other amino acid recoveries may be compromised because of complete or partial destruction (e.g., tryptophan, cysteine, threonine, serine, methionine), incomplete bond cleavage (i.e., for isoleucine and valine), and free amino acid contamination (i.e., by glycine and serine).
Because those amino acids that are recovered best represent the protein, these amino acids are chosen to quantify the amount of protein. Well-recovered amino acids are, typically, aspartate-asparagine, glutamate-glutamine, alanine, leucine, phenylalanine, lysine, and arginine. This list can be modified based on experience with one's own analysis system. Divide the quantity, in nmol, of each of the well-recovered amino acids by the expected number of residues for that amino acid to obtain the protein content based on each well-recovered amino acid. Average the protein content results calculated. The protein content determined for each of the well-recovered amino acids should be evenly distributed about the mean. Discard protein content values for those amino acids that have an unacceptable deviation from the mean. Typically, a greater than 5% variation from the mean is considered unacceptable, but this is arbitrary. Recalculate the mean protein content from the remaining values to obtain the protein content of the sample. Divide the content of each amino acid by the calculated mean protein content to determine the amino acid composition of the sample by analysis.
Calculate the relative compositional error, in percentage, by the formula:
100m / mS,
in which m is the experimentally determined quantity, in nmol per amino acid residue, of the amino acid under test; and mS is the known residue value for that amino acid. The average relative compositional error is the average of the absolute values of the relative compositional errors of the individual amino acids, typically excluding tryptophan and cysteine from this calculation. The average relative compositional error can provide important information on the stability of analysis run over time. The agreement in the amino acid composition between the protein sample and the known composition can be used to corroborate the identity and purity of the protein in the sample.

CAPILLARY ELECTROPHORESIS
Capillary electrophoresis is a physical method of analysis based on the migration, inside a capillary, of charged analytes dissolved in an electrolyte solution, under the influence of a direct-current electric field. In this section we are describing four capillary electrophoresis methods, Free Solution Capillary Electrophoresis, Capillary Gel Electrophoresis, Capillary Isoelectric Focusing, and Micelle Electrokinetic Chromatography.
General Principle
The migration velocity of the analyte under an electric field of intensity, E, is determined by the electrophoretic mobility of the analyte and the electroosmotic mobility of the buffer inside the capillary. The electrophoretic mobility of a solute (µep) depends on the characteristics of the solute (electrical charge, molecular size, and shape) and the characteristics of the buffer in which the migration takes place (type and ionic strength of the electrolyte, pH, viscosity, and additives). The electrophoretic velocity (Vep) of a solute, assuming a spherical shape, is as follows:
Click to View Image
in which q is the effective charge of the particle; is the viscosity of the buffer; r is the size of the solute ion; V is the applied voltage; and L is the total length of the capillary.
When an electric field is applied through the capillary filled with buffer, a flow of solvent is generated inside the capillary called electroosmotic flow. Its velocity depends on the electroosmotic mobility (µeo) which in turn depends on the charge density on the capillary internal wall and the buffer characteristics. The electroosmotic velocity (Veo) is as follows:
Click to View Image
in which is the dielectric constant of the buffer; is the zeta potential of the capillary surface; and the other terms are as defined above.
The electrophoretic and electroosmotic mobilities of the analyte may act in the same direction or in opposite directions, depending on the charge (positive or negative) of the solute, and the velocity of the solute (v) is as follows:
V = Vep ± Veo.
The sum or the difference between the two velocities (Vep and Veo) is used depending on whether the mobilities act in the same or opposite directions. Under conditions with a fast Veo, with respect to the Vep of the solutes, both negative and positive charged analytes can be separated in the same run. The time (t) taken by the solute to migrate the distance (l) from the injection end of the capillary to the detection point (capillary effective length) is as follows:
Click to View Image
in which the other terms are as defined above.
In general, the fused-silica capillaries used in electrophoresis bear negative charges on the inner wall, producing electroosmotic flow towards the cathode. The electroosmotic flow has to remain constant from run to run to obtain good reproducibility in the migration velocity of the solutes. For some applications, it might be necessary to reduce or suppress the electroosmotic flow by modifying the inner wall of the capillary or by changing the pH of the buffer solution.
When the sample is introduced in the capillary, each analyte ion of the sample migrates within the background electrolyte as an independent zone according to its electrophoretic mobility. The spreading of each solute band (zone dispersion) results from a different phenomena. Under ideal conditions, the sole contribution to the solute-zone broadening is molecular diffusion of the solute along the capillary (longitudinal diffusion). In this case, the efficiency of the zone is expressed as the number of theoretical plates (N), as follows:
Click to View Image
in which D is the molecular diffusion of the solute in the buffer; and the other terms are as defined above.
From a practical point of view, other phenomena such as heat dissipation, sample adsorption onto the capillary wall, mismatched conductivity between sample and buffer, length of the injection plug, detector cell size, and unleveled buffer reservoirs can also significantly contribute to band dispersion. Separation between two bands (expressed by the resolution, RS) can be achieved by modification of the electrophoretic mobility of the analytes, by the electroosmotic mobility induced by capillary, and by increasing the efficiency for the band of each analyte as follows:
Click to View Image
in which µepa and µepb are the electrophoretic mobilities of the two compounds to be separated; bar(µ)ep is the average electrophoretic mobility of the two solutes calculated as:
bar(µ)ep = ½ (µepb + µepa),
and the other terms are as defined above.
Apparatus
An apparatus for capillary electrophoresis is composed of a high voltage controllable power supply; two buffer reservoirs held at the same level and containing specified anodic and cathodic solutions; two electrodes assemblies (cathode and anode) immersed in the buffer reservoirs and connected to the power supply; a separation capillary usually made of fused-silica, with sometimes an optical viewing window aligned with detector, depending on the detector, with the ends of the capillary placed in the buffer reservoirs and the capillary being filled with a solution specified in a given monograph; a suitable injection system; a detector capable of monitoring the amount of substance of interest passing through a segment of the separation capillary at a given time, generally based on absorption spectrophotometry (UV and visible), fluorimetry, conductimetric, amperometric, or mass spectrometric detection, depending on the specific applications, or even indirect detection to detect non-UV-absorbing and nonfluorescent compounds; and a thermostatic system capable of maintaining the temperature inside the capillary.
The method of injection of samples and its automation is critical for precise quantitative analysis. Methods of injection include gravity, pressure or vacuum, or electrokinetic injection. The amount of each sample component introduced electrokinetically depends on its electrophoretic mobility, thus possibly biasing the results.
It is expected that the capillary, the buffer solutions, the preconditioning method, the sample solution, and the migration conditions will be specified in the individual monograph. The electrolytic solution employed may be filtered to remove particles and degassed to avoid bubble formation that could interfere with the detection system. To achieve reproducible migration time of the solutes, if would be necessary to develop, for each analytical method, a rigorous rinsing routine after each injection.
Free Solution Capillary Electrophoresis
In free solution capillary electrophoresis, analytes are separated in a capillary containing only buffer without any anticonvective medium. In this technique, separation takes place because the different components of the sample migrate as discrete bands with different velocities. The velocity of each band depends on the electrophoretic mobility of the solute and the electroosmotic flow on the capillary. Coated capillaries, with reduced electroosmotic flow, can be used to increase the separation capacity of those substances absorbing on fused silica surfaces.
This mode of capillary electrophoresis is appropriate for the analysis of small (MW < 2000) and large molecules (2000 < MW < 100,000). Due to the high efficiency achieved, molecules having only minute differences in their charge-to-mass ratio can be separated. This method also allows the separation of chiral compounds by adding chiral selectors to the separation buffer. The optimization of the separations requires consideration of a number of instrumental and electrolytic solution parameters.
INSTRUMENTAL PARAMETERS
Voltage— The separation time is universally proportional to applied voltage. However, an increase in the voltage used can cause excessive heat production, giving rise to temperature and viscosity gradients in the buffer inside the capillary, which causes band broadening and decreases resolution.
Temperature— The main effect of temperature is observed on buffer viscosity and electrical conductivity, thus affecting migration velocity. In some cases, an increase in capillary temperature can cause a conformational change of some proteins, modifying their migration time and the efficiency of the separation.
Capillary— The length and internal diameter of the capillary affects the analysis time, the efficiency of separations, and the load capacity. Increasing both effective length and total length can decrease the electric fields, at a constant voltage, which will increase migration time. For a given buffer and electric field, heat dissipation (thus sample band broadening) depends on the internal diameter of the capillary. The latter also affects the detection limit, depending on the sample volume injected into the capillary and the detection system used.
The adsorption of sample components on the capillary wall limits efficiency; therefore, methods to avoid these interactions should be considered in the development of a separation method. This is critical in samples containing proteins. Strategies have been devised to avoid adsorption of proteins on the capillary wall. These strategies include both the use of extreme pH and the absorption of positively charged buffer additives that only require modification of the buffer composition. Other strategies include the coating of the internal wall of the capillary with a polymer covalently bonded to the silica that prevents interaction between the proteins and the negatively charged silica surface. Capillaries with coatings consisting of neutral-hydrophilic, cationic, and anionic polymers are commercially available.
ELECTROLYTIC SOLUTION PARAMETERS
Buffer Type and Concentrations— Suitable buffers for capillary electrophoresis have an appropriate buffer capacity in the pH range of choice and low mobility to minimize current generation.
To minimize peak shape distortion, it is important to match buffer–ion mobility to solute mobility, whenever possible. The type of sample solvent used is important to achieve on-column sample focusing which increases separation efficiency and improves detection. Also, an increase in buffer concentration at a given pH will decrease electroosmotic flow and solute velocity.
Buffer pH— The pH of the buffer can affect separation by modifying the charge of the analyte or other additives and by changing the electroosmotic flow. For protein and peptide separation, a change in the pH of the buffer from above the isoelectric point to below the isoelectric point changes the net charge of the solute from negative to positive. An increase in the buffer pH generally increases the electroosmotic flow.
Organic Solvents— Organic modifiers, such as methanol, acetonitrile, and others, are added to the aqueous buffer to increase the solubility of the solute or other additives and/or to affect the ionization degree of the sample components. The addition of these organic modifiers to the buffer generally causes a decrease in the electroosmotic flow.
Additives for Chiral Separations— To separate optical isomers, a chiral selector is added to the separation buffer. The most commonly used chiral selectors are cyclodextrins, although in some cases crown ethers, certain polysaccharides, or even proteins can be used. Because chiral recognition is governed by the different interactions between the chiral selector and each of the enantiomers, the resolution achieved for the chiral compounds depends largely on the type of chiral selector used. While developing a given separation it may be useful to test cyclodextrins having a different cavity size (-, -, or G-cyclodextrin) or modified cyclodextrins with neutral (methyl, ethyl, hydroxyalkyl, etc.) or ionizable (aminomethyl, carboxymethyl, sulfobutylether, etc.) moities. The resolution of chiral separations is also controlled by the concentration of the chiral selector, the composition and pH of the buffer, and the separation temperature. Organic additives, such as methanol or urea, can also affect the resolution of separation.
Capillary Gel Electrophoresis
Separation takes place inside a capillary filled with a polymer acting as a molecular sieve. The smaller components in the sample move faster along the capillary than the larger ones. This method can be used for separation of biopolymers-proteins, and DNA fragments, according to their molecular mass.
CHARACTERISTICS OF CHEMICAL and PHYSICAL GELS
Chemical Gels— Chemical gels are prepared inside the capillary by reaction of monomers. One example of such a gel is a cross-linked polyacrylamide. This type of gel is bonded to the fused-silica wall and cannot be removed without destroying the capillary. For protein analysis, the separation buffer usually contains sodium dodecyl sulfate and the sample is denatured by heating in a mixture of sodium dodecyl sulfate and 2-mercaptoethanol or dithiothreitol before injection. Optimization of separation in a cross-linked gel is obtained by modifying the separation buffer (see Free Solution Capillary Electrophoresis) and by controlling the gel porosity during the gel preparation. For a cross-linked polyacrylamide gel, the porosity can be modified by changing the concentration of acrylamide and/or the ratio of the cross-linker. As a rule, a decrease in the porosity of the gel leads to a decrease in the mobility of the solutes. Due to the rigidity of this type of gel, only electrokinetic injection can be used.
Physical Gels— Physical gels are hydrophilic polymers (i.e., linear polyacrylamide, cellulose derivatives, dextran, etc.) which can be dissolved in aqueous separation buffers, giving rise to a separation medium that also acts as a molecular sieve. These polymeric separation media are easier to prepare than cross-linked polymers. They can be prepared in a vial and filled by pressure in a wall-coated capillary with no electroosmotic flow. Replacing the gel before every injection generally improves the separation reproducibility. The porosity of the physical gels can be increased by using polymers of higher molecular weight (at a given polymer concentration) or by decreasing the polymer concentration (for a given polymer molecular weight). A decrease in gel porosity leads to a decrease in the mobility of the solute for the same buffer. Both hydrodynamic and electromigration injection techniques can be used, since the dissolution of these polymers in the buffer gives low viscosity solutions.
Capillary Isoelectric Focusing
The molecules migrate under the influence of the electric field, so long as they are charged, in a pH gradient generated by ampholytes having pI values in a wide range (poly-aminocarboxylic acids) dissolved in the separation buffer. The three basic steps in capillary isoelectric focusing are loading, focusing, and mobilization.
Loading—
Loading in One Step— The sample is mixed with ampholytes and introduced into the capillary by pressure or vacuum.
Sequential Loading— A leading buffer, then the ampholytes, then the sample mixed with ampholytes, again ampholytes alone, and finally the terminating buffer are introduced into the capillary. The volume of the sample must be small enough so as to not modify the pH gradient.
Focusing— When the voltage is applied, ampholytes migrate toward the cathode or the anode according to their net charge, creating the pH gradient from anode (lower pH) to cathode (higher pH). The components to be separated migrate until they reach a pH corresponding to their isoelectric point and the current drops to very low values.
Mobilization— The bands of separated components migrate past the detector by one of the three following methods.
Method 1— During Focusing, under the influence of the electroosmotic flow when this flow is small enough to allow the focusing of the components.
Method 2— By application of positive pressure after Focusing.
Method 3— After Focusing, by adding salts in the cathode reservoir or the anode reservoir (depending on the direction chosen for mobilization), in order to alter the pH in the capillary when the voltage is applied. As the pH is changed, the proteins and ampholytes are mobilized in the direction of the reservoir which contains added salts and pass the detector.
The separation achieved is expressed as DpI and depends on the pH gradient (dpH), the number of ampholytes having different pI values, the diffusion coefficient (D), the intensity of the electric field (E), and the variation of the electrophoretic mobility of the analyte with the pH, and is as follows:
Click to View Image
in which dpH/dx is the pH gradient; and –dµ/dpH is the variation of the solution mobility with the pH in the region close to the pI.
Optimization Parameters— The major parameters that need to be considered in the development of separations are voltage, capillary, and solutes.
Voltage— Use of high fields from 300 V/cm to 1,000 V/cm during Focusing.
Capillary— Depending on the Mobilization strategy selected (see above), the electroosmotic flow must be reduced or suppressed. Coated capillaries tend to reduce the electroosmotic flow.
Solutions— The anode buffer reservoir is filled with a solution of a lower pH than the pI of the most acidic ampholyte and the cathode reservoir is filled with a solution with a higher pH than the pI of the most basic ampholyte. Phosphoric acid for the anode and sodium hydroxide for the cathode are frequently used.
Addition of a polymer, like methylcellulose, in the ampholyte solution tends to suppress convective forces (if any) and electroosmotic flow by increasing the viscosity. Commercial ampholytes covering many pH ranges are available and may also be mixed to obtain an expanded pH range. Broad pH ranges are used to estimate the isoelectric point whereas narrower ranges are employed to improve accuracy. Calibration can be made by correlating migration time with the isoelectric point of a series of standard protein markers. During Focusing, precipitation of proteins at their isoelectric point can be prevented, if necessary, using buffer additives such as glycerol, surfactants, urea, or Zwitterionic buffers. However, depending on the concentration, urea can denature proteins.
Micellar Electrokinetic Chromatography (MEKC)
Separation takes place in an electrolytic solution which contains a surfactant, generally ionic, at a concentration above the critical micellar concentration. The solute molecules are distributed between the aqueous buffer and the pseudo-stationary phase composed by the micelles according to the solute's partition coefficient. The technique can be considered as a hybrid of electrophoresis and chromatography. It is an electrophoretic technique that can be used for the separation of both neutral and charged solutes maintaining the efficiency, speed, and instrumental suitability of capillary electrophoresis. One of the most widely used surfactants is sodium dodecyl sulfate, although other anionic and cationic surfactants, such as cetyl trimethyl ammonium salts, have also been used.
At neutral and alkaline pH, a strong electroosmotic flow is generated and moves the separation buffer ions in the direction of the cathode. If sodium dodecyl sulfate is used as surfactant, the electrophoretic migration of the anionic micelle is in the opposite direction, toward the anode. As a result, the overall micelle migration velocity is slowed compared to the bulk flow of the electrolytic solution. In the case of neutral solutes, since the analyte can partition between the micelle and the aqueous buffer and has no electrophoretic mobility, the analyte migration velocity will only depend on the partition coefficient between the micelle and the aqueous buffer. In the electrophoretogram, the peak corresponding to each uncharged solute is always between that of the electroosmotic flow marker and that of the micelle, and the time elapsed between these two peaks is called the separation window. For electrically charged solutes, the migration velocity depends on both the partition coefficient of the solute between the micelle and the aqueous buffer and on the electrophoretic mobility of the solute in the absence of micelles.
The separation mechanism is essentially chromatographic, and migration of the solute and resolution can be expressed in terms of the capacity factor of the solute (K¢), which is the ratio between the total number of moles of solute in the micelle to those in the mobile phase. For a neutral compound, K¢ is as follows:
Click to View Image
in which tr is the migration time of the solute; to is the analysis time of the unretained solute obtained by injecting an electroosmotic flow marker which does not enter the micelle (i.e., methanol); tm is the micelle migration time measured by injecting a micelle marker, such as Sudan III, which migrates continuously associated in the micelle; K is the partition coefficient of the solute; VS is the volume of the micelles phase; and VM is the volume of the mobile phase.
The resolution between two closely-migrating compounds (RS) is as follows:
Click to View Image
in which N is the number of theoretical plates for one of the compounds; is the selectivity obtained; K¢a and K¢b are capacity factors for both components; and the other terms are as defined above.
Similar, but not identical, equations give K¢ and RS values for electrically charged compounds.
Optimization Parameters— The main parameters to be considered in the development of separations by MEKC are instrumental and electrolytic solution parameters.
INSTRUMENTAL PARAMETERS
Voltage— Separation time is inversely proportional to applied voltage. An increase in voltage can cause excessive heat production that gives rise to temperature gradients and viscosity gradients of the buffer in the cross section of the capillary. This effect can be significant with high conductivity buffers, such as those containing micelles. Poor heat dissipation causes band broadening and decreases resolution.
Temperature— Variations in capillary temperature affect the partition coefficient of the solute between the buffer and the micelle, the critical micelle concentration, and the viscosity of the buffer. These parameters contribute to the migration time of the solutes.
Capillary— Length and internal diameter contribute to analysis time and efficiency of separations. Increasing both effective length and total length can decrease the electrical fields, working at constant voltage, and will increase migration time and improve the separation efficiency. The internal diameter controls heat dissipation, at a given buffer and electrical field, and provides a broadening of the sample band.
ELECTROLYTIC SOLUTION PARAMETERS
Surfactant Type and Concentration— The type of surfactant, as the stationary phase in chromatography, affects the resolution since it modifies separation selectively. The log K¢ of a neutral compound increases linearly with the concentration of detergent in the mobile phase. Resolution in MEKC reaches a maximum when K¢ approaches the value of
Click to View Image
modifying the concentration of surfactant in the mobile phase changes the resolution.
Buffer pH— pH does not modify the partition coefficient of non-ionized solutes, but it can modify the electroosmotic flow in uncoated capillaries. A decrease in the buffer pH decreases the electroosmotic flow and therefore increases the resolution of the neutral solutes, giving rise to longer analysis time.
Organic Solvents— To improve separation of hydrophobic compounds, organic modifiers (methanol, propanol, acetonitrile, etc.) can be added to the separation electrolytic solution. The addition of these modifiers generally decreases migration time and selectivity of the separation. The addition of organic modifiers affects micelle formation, thus a given surfactant concentration can be used only with a certain percentage of organic modifier before the micellezation equilibrium is eliminated or adversely affected, resulting in the absence of micelles and therefore the absence of the partition mechanism of MEKC. The elimination of micelles in the presence of a high content of organic solvent does not always mean that the separation will no longer be possible, since in some cases, the hydrophobic interaction between the ionic surfactant monomer and the neutral solutes form solvophobic complexes that can be separated electrophoretically.
Additives for Chiral Separations— A chiral selector is included in the micellar system, either covalently bound to the surfactant or added to the micellar separation electrolyte. Micelles which have a moiety with chiral discrimination properties include salts, N-dodecanoyl-L-amino acids, bile salts, etc. Chiral resolution can also be achieved using chiral discriminators, such as cyclodextrins added to the electrolytic solutions which contain micelliced achiral surfactants.
Other Additives— Selectivity can be modified by adding chemicals to the buffer. Addition of several types of cyclodextrins to the buffer are also used to reduce the interaction of hydrophobic solutes with the micelle, increasing the selectivity for this type of compound. The addition of substances able to modify solute-micelle interactions by adsorption on the latter, has been used to improve the selectivity of the separations in MEKC. These additives may consist of a second surfactant (ionic or non-ionic) which gives rise to mixed micelles, metallic cations which dissolve in the micelle, and give co-ordination complexes with the solutes.
QUANTITATIVE ANALYSIS
Peak areas must be divided by the corresponding migration time to give the corrected area in order to compensate for the shift in migration time from run to run, thus reducing the variation of the response. It will also compensate for the different responses of sample constituents with different migration times. Where an internal standard is used, check that no peak of the substance to be examined is masked by that of the internal standard.
Calculations— From the values obtained, calculate the content of a component or components being determined. When indicated, the percentage of one (or more) components of the sample to be examined is calculated by determining the areas of the peak(s) as a percentage of the total corrected areas of all the peaks, excluding those due to solvents or any added reagents. The use of an automatic integration system (integrator or data acquisition and processing system) is recommended.
Capillary Electrophoresis System Suitability
The choice of suitability parameters to be used will depend on the type of capillary electrophoresis that is performed. These parameters are the capacity factor (K¢) used only for Micelles Electrokinetic Chromatography, the number of theoretical plates (n), the symmetry factor (AS), and the resolution (RS). Note that in previous sections, the theoretical expression for n and RS have been described, but more practical equations that allow for the determination of these suitability parameters using the electrophoretograms are described below.
The number of theoretical plates (n) may be calculated from the formula:
n = 5.54 (t / b0.5)2,
in which t is the distance, in mm, along the baseline between the point of injection and the perpendicular dropped from the maximum of the peak in question; and b0.5 is the peak width, in mm, at half height.
The resolution (RS) may be calculated from the formula:
RS = 1.18(tb ta / b0.5b + b0.5a),
in which tb and ta are the distances, in mm, along the baseline, between the point of injection and the perpendicular dropped from the maxima of two adjacent peaks (tb > ta); and b0.5b and b0.5a are the peak widths, in mm, at half height.
The resolution (RS) may be also calculated by measuring the height of the valley (c) between two partly resolved peaks in a standard preparation, the height of the smaller peak (d), and by specifying (c/d)x, in which x is the limit indicated in the individual monograph.
The symmetry factor of a peak (AS) may be calculated using the formula:
AS = b0.05/2A,
in which b0.05 is the width of the peak at one-twentieth of the peak height; and A is the distance between the perpendicular dropped from the peak maximum and the leading edge of the peak at one-twentieth of the peak height.
Other system suitability parameters include tests for area repeatability (i.e., standard deviation of areas or of area/migration time) and tests for migration time repeatability (i.e., standard deviation of migration time). For migration time repeatability, it will be necessary to provide for a test to measure the suitability of the capillary washing procedures. To avoid the lack of repeatability of the migration time, an alternative practice is to use a migration time relative to an internal standard.
A test for the verification of the signal-to-noise ratio for a standard preparation or the determination of the limit of quantitation is a useful system suitability parameter. The detection limit and quantitation limit correspond to a signal-to-noise ratio greater than 3 and 10, respectively. The signal-to-noise ratio (S/N) is calculated as follows:
S/N = 2H/hn,
in which H is the height of the peak corresponding to the component concerned in the electrophoretogram obtained with the specified reference solution; and hn is the absolute value of the largest noise fluctuation from the baseline in an electrophoretogram obtained after injection of a blank and observed over a distance equal to twenty times the width at the half-height of the peak in the electrophoretogram obtained with the reference solution, and situated equally around the place where this peak would be found.

ISOELECTRIC FOCUSING
Isoelectric focusing (IEF) is a method of electrophoresis that separates proteins according to their isoelectric points. Separation is carried out in a slab of polyacrylamide or agarose gel that contains a mixture of amphoteric electrolytes (ampholytes). When subjected to an electrical field, the ampholytes migrate in the gel to create a pH gradient. In some cases, gels containing an immobilized pH gradient, prepared by incorporating weak acids and bases to specific regions of the gel network during the preparation of the gel, are used. When the applied proteins reach the gel fraction that has a pH that is the same as their isoelectric point, their charge is neutralized and migration ceases. Gradients can be made over various ranges of pH, according to the mixture of ampholytes chosen.
General Principles
When a protein is at the position of its isoelectric point, it has no net charge and cannot be moved in a gel matrix by the electric field. It may, however, move from that position by diffusion. The pH gradient forces a protein to remain in its isoelectric point position, thus concentrating it; this concentration effect is called ``focusing''. Increasing the applied voltage or reducing the sample load results in improved resolution of bands. The applied voltage is limited by the heat generated because the heat must be dissipated. The use of thin gels and an efficient cooling plate controlled by a thermostatic circulator prevents the burning of the gel while allowing sharp focusing. The separation is estimated by determining the minimum pI difference, which is necessary to separate two neighboring bands, as follows:
Click to View Image
in which D is the diffusion coefficient of the protein; dpH/dx is the pH gradient; E is the intensity of the electric field, in volts per centimeter; and –dµ/dpH is the variation of the solute mobility with the pH in the region close to the pI. Since D and –dµ/dpH for a given protein cannot be altered, the separation can be improved by using a narrower pH range and by increasing the intensity of the electric field.
From an operational point, special attention must be paid to sample characteristics and/or preparation. Salt in a sample can be problematic and it is best to prepare the sample, if possible, in deionized water or 2% ampholytes using dialysis or gel filtration if necessary. Potentials of 2500 volts have been used and are considered optimal under a given set of conditions. Up to 30 watts of constant power can be applied and will generally give complete separation in 1.5 to 3.0 hours. The time required for completion of focusing in thin-layer polyacrylamide gels is determined by placing a colored protein (e.g., hemoglobin) at different positions on the gel surface and by applying the electric field: the steady state is reached when all applications give an identical band pattern. In some procedures the completion of the focusing is indicated by the time elapsed after the sample application.
Resolution between protein bands on an IEF gel prepared with carrier ampholytes can be quite good. Better resolution may be achieved by using immobilized pH gradients where the buffering species, which are analogous to carrier ampholytes, are copolymerized within the gel matrix. Proteins exhibiting pls differing by as little as 0.02 pH units may be resolved using a gel prepared with carrier ampholytes, while immobilized pH gradients can resolve protein differing by approximately 0.001 pH units.
The IEF gel can be used as an identity test when migration on the gel is compared to a standard preparation and IEF calibration proteins, the IEF gel can be used as a limit test when the density of a band on IEF is compared subjectively with the density of bands appearing in a standard preparation, or it can be used as a semi-quantitative test when the density is measured using a densitometer or similar instrumentation to determine the relative concentration of protein in the bands.
Apparatus
An apparatus for isoelectric focusing consists of a controllable direct current generator, of stabilized output; a rigid plastic isoelectric focusing chamber that contains a cooled plate of suitable material to support the gel; and a plastic cover with platinum electrodes that are connected to the gel by means of paper wicks of suitable width, length, and thickness, impregnated with solutions of anodic and cathodic electrolytes.
Procedure
Unless otherwise indicated in a given monograph, the following procedure in thick polyacrylamide slab gels is to be used.
Preparation of the Gels—
Assembly— Composed of a glass plate (A) on which a polyester film (B) is placed to facilitate handling of the gel, one or more spacers (C), a second glass plate (D), and clamps to hold the structure together (see Figure 1).
Click to View Image
7.5% Polyacrylamide Gel— Dissolve 29.1 g of acrylamide and 0.9 g of methylenebisacrylamide in 100 mL of water. To 2.5 volumes of this solution, add the mixture of ampholytes specified in the individual monograph, and make up to 10 volumes with water. Mix carefully, and degas the solution.
Preparation of the Assembly— Place the polyester film on the lower glass plate, apply the spacer, place the second glass plate, and fit the clamps. Before use, place the mixture on a magnetic stirrer, and add 0.25 volumes of a 10% solution of ammonium persulfate and 0.25 volumes of tetramethylenediamine. Immediately fill the space between the glass plates of the assembly with the gel.
Fixing Solution for Isoelectric Focusing Polyacrylamide Gel— Mix 35 g of sulfosalicylic acid and 100 g of trichloroacetic acid in 1000 mL of water.
Coomassie Staining Solution and Destaining Solution— Use the same solutions indicated in Polyacrylamide Gel Electrophoresis.
Procedure— Dismantle the assembly, and using the polyester film, transfer the gel onto the cooled support wetted with a few milliliters of a suitable liquid, taking care to avoid forming air bubbles. Prepare the test solutions and reference solutions as specified in the individual monograph. Place strips of paper for sample application, about 10 mm × 5 mm in size, on the gel, and impregnate each with the prescribed amount of the test and reference solutions. If the protein concentration of the solution is too low, several strips may be superimposed (up to four). Also apply the prescribed quantity of a solution of proteins with known isoelectric points as pH markers to calibrate the gel. In some procedures, the gel has pre-cast slots where a solution of the sample is applied instead of using impregnated paper strips. Cut two strips of paper to the length of the gel, and impregnate them with the electrolyte solutions: acid for the anode and alkaline for the cathode. The compositions of the anode and cathode solutions are given in the individual monograph. Apply these paper wicks to each side of the gel several millimeters from the edge. Fit the cover so that the electrodes are in contact with the wicks (with respect to the anodic and cathodic poles). Proceed with the isoelectric focusing by applying the electrical parameters described in the individual monograph. Switch off the current when the migration of the mixture of standard proteins has stabilized. Using forceps, remove the sample application strips and the two electrode wicks. Immerse the gel in Fixing Solution for Isoelectric Focusing Polyacrylamide Gel. Incubate with gentle shaking at room temperature for 30 minutes. Drain off the solution, and add 200 mL of Destaining Solution. Incubate with shaking for 1 hour. Drain the gel, add Coomassie Staining Solution. Incubate for 30 minutes. Destain the gel by passive diffusion with Destaining Solution until the bands are well visualized against a clear background. Locate the position and intensity of the bands in the electropherogram as prescribed in the individual monograph.
Alternative Procedure— When a monograph references the general method for isoelectric focusing above, variations in methodology or procedure may be used, subject to validation. These variations include the use of commercially available pre-cast gels; the use of immobilized pH gradients; the use of rod gels; and the use of cassettes of different dimensions, including ultra-thin (0.2 mm) gels; the variations in the sample application procedure, including different sample volumes or the use of sample application masks or wicks other than paper; the use of alternate running conditions, including variations in the electric field depending on gel dimensions and equipment, and the use of fixed migration times rather than subjective interpretation of band stability; the inclusion of a pre-focusing step; the use of automated instrumentation; and the use of agarose gels.
Validation of Procedure
Where alternative methods to the general method are employed, they must be validated. The following criteria may be used to validate the separation: the formation of a stable pH gradient of desired characteristics, evaluated using colored pH markers of known isoelectric points; the comparison with electropherogram provided with the chemical reference substance for the preparation to be examined; and any other validation criteria as prescribed in the individual monograph.
Specified Variations to the General Method
Variations to the general method required for the analysis of specific substances may be specified in detail in individual monographs. Variations may include the addition of urea in the running gel (3 M concentration is often satisfactory to keep protein in solution but up to 8 M can be used). Some proteins precipitate at their isoelectric point. In this case, urea is included in the gel formulation to keep the protein in solution. If urea is used, only fresh solutions should be used to prevent carbamylation of the protein. Other variations include the use of alternative staining methods and the use of gel additives such as non-ionic detergents (e.g., octylglucoside) or zwitterionic detergents (e.g., CHAPS or CHAPSO) to prevent proteins from aggregating or precipitating.
NOTE—The following are general precautionary items that can be used to improve the method. Samples can be applied to any area on the gel, but in general, they should be applied to areas where they are expected to focus. To protect the proteins from extreme pH environments, samples should not be applied close to either electrode. During method development, the analyst can try applying the protein in three positions on the gel (i.e., middle and both ends); the pattern of a protein applied at opposite ends of the gel may not be identical. A phenomenon known as cathodic drift, where the pH gradient decays over time, may occur if a gel is focused too long. Although not well understood, electroendoosmosis and absorption of carbon dioxide may be factors that lead to cathodic drift. Cathodic drift is observed as focused protein migrating off the cathode end of the gel. Immobilized pH gradients may be used to address this problem. Efficient cooling (approximately 4) of the bed that the gel lies on during focusing is important. High field strengths used during isoelectric focusing can lead to overheating and affect the quality of the focused gel.

PEPTIDE MAPPING
Purpose and Scope
Peptide mapping is an identity test for proteins, especially those obtained by r-DNA technology. It involves the chemical or enzymatic treatment of a protein resulting in the formation of peptide fragments followed by separation and identification of the resultant fragments in a reproducible manner. It is a powerful test that is capable of identifying single amino acid changes resulting from events such as errors in the reading of complementary DNA (cDNA) sequences or point mutations. Peptide mapping is a comparative procedure because the information obtained, compared to a reference standard or reference material similarly treated, confirms the primary structure of the protein, is capable of detecting whether alterations in structure have occurred, and demonstrates process consistency and genetic stability. Each protein presents unique characteristics which must be well understood so that the scientific and analytical approaches permit validated development of a peptide map that provides sufficient specificity.
This section provides detailed assistance in the application of peptide mapping and its validation to characterize the desired protein product, to evaluate the stability of the expression construct of cells used for recombinant DNA products, to evaluate the consistency of the overall process, and to assess product stability, as well as to ensure the identity of the protein product, or to detect the presence of protein variant. The validation scheme presented differentiates between qualification of the method at an early stage in the regulatory process, Investigational New Drug (IND) level, and full validation in support of New Drug Application (NDA), Product License Application (PLA), or Marketing Authorization Application (MAA). The validation concepts described are consistent with the general information chapter Validation of Compendial Methods 1225 and with the International Conference on Harmonization (ICH) document on Analytical Methods Validation.
The Peptide Map
Peptide mapping is not a general method, but involves developing specific maps for each unique protein. Although the technology is evolving rapidly, there are certain methods that are generally accepted. Variations of these methods will be indicated, when appropriate, in specific monographs.
A peptide map may be viewed as a fingerprint of a protein and is the end product of several chemical processes that provide a comprehensive understanding of the protein being analyzed. Four major steps are necessary for the development of the procedure: isolation and purification of the protein, if the protein is part of a formulation; selective cleavage of the peptide bands; chromatographic separation of the peptides; and analysis and identification of the peptides. A test sample is digested and assayed in parallel with a reference standard or reference material. Complete cleavage is more likely to occur when enzymes such as endoproteases (e.g., trypsin) are used instead of chemical cleavage reagents. A map should contain enough peptides to be meaningful. On the other hand, if there are too many fragments, the map might lose its specificity because many proteins will then have the same profiles.
ISOLATION and PURIFICATION
Isolation and purification are necessary for analysis of bulk drugs or dosage forms containing interfering excipients and carrier proteins and, when required, will be specified in the monograph. Quantitative recovery of protein from the dosage form should be validated.
SELECTIVE CLEAVAGE OF PEPTIDE BONDS
The selection of the approach used for the cleavage of peptide bonds will depend on the protein under test. This selection process involves determination of the type of cleavage to be employed—enzymatic or chemical—and the type of cleavage agent within the chosen category. Several cleavage agents and their specificity are shown in Table 1.
Table 1. Examples of Cleaving Agents
Type Agent Specificity
Enzymatic Trypsin (EC 3.4.21.4) C-terminal side of Arg and Lys
Chymrotrypsin (EC 3.4.21.1)
C-terminal side of hydrophobic residues
(e.g., Leu, Met, Ala, aromatics)
Pepsin (EC 3.4.23.122) Nonspecific digest
Lysyl endopeptidase (Lys-C Endopeptidase) C-terminal side of Lys
(EC 3.4.21.50)
Glutamyl endopeptidase (from S. aureus stain V8) C-terminal side of Glu and Asp
(EC 23.4.21.19)
Peptidyl-Asp metaplo-endopeptidase N-terminal side of Asp
(Endoproteinase Asp-N)
(EC 3.4.24.33)
(Clostripan) C-terminal side of Arg
(EC 3.4.28.8)
Chemical Cyanogen bromide C-terminal side of Met
2-Nitro-5-thio-cyano- benzoic acid N-terminal side of Cys
O-Iodosobenzoic acid C-terminal side of Trp and Tyr
Dilute acid Asp and Pro
BNPS-skatole Trp
This list is not all-inclusive and will be expanded as other cleavage agents are identified.
Pretreatment of Sample— Depending on the size or the configuration of the protein, different approaches in the pretreatment of samples can be used. For monoclonal antibodies, the heavy and light chains will need to be separated before mapping. If trypsin is used as a cleavage agent for proteins with a molecular mass greater than 100,000 Da, lysine residues must be protected by citraconylation or maleylation; otherwise, too many peptides will be generated.
Pretreatment of the Cleavage Agent— Pretreatment of cleavage agents—especially enzymatic agents—might be necessary for purification purposes to ensure reproducibility of the map. For example, trypsin used as a cleavage agent will have to be treated with tosyl-L-phenylalanine chloromethyl ketone to inactivate chymotrypsin. Other methods, such as purification of trypsin by HPLC or immobilization of enzyme on a gel support, have been successfully used when only a small amount of protein is available.
Pretreatment of the Protein— Under certain conditions, it might be necessary to concentrate the sample or to separate the protein from added substances and stabilizers used in the formulation of the product, if these interfere with the mapping procedure. Physical procedures used for pretreatment can include ultrafiltration, column chromatography, and lyophilization.
Other pretreatments, such as the addition of chaotropic agents (e.g., urea), can be used to unfold the protein prior to mapping. To allow the enzyme to have full access to cleavage sites and permit some unfolding of the protein, it is often necessary to reduce and alkylate the disulfide bonds prior to digestion.
Digestion with trypsin can introduce ambiguities in the tryptic map due to side reactions occurring during the digestion reaction, such as nonspecific cleavage, deamidation, disulfide isomerization, oxidation of methionine residues, or formation of pyroglutamic groups created from the deamidation of glutamine at the N-terminal side of a peptide. Furthermore, peaks may be produced by autohydrolysis of trypsin. Their intensities depend on the ratio of trypsin to protein. To avoid autohydrolysis, solutions of proteases may be prepared at a pH that is not optimal (e.g., at pH 5 for trypsin), which would mean that the enzyme would not become active until diluted with the digest buffer.
Establishment of Optimal Digestion Conditions— Factors that affect the completeness and effectiveness of digestion of proteins are those that could affect any chemical or enzymatic reactions.
pH— The pH of the digestion mixture is empirically determined to ensure the optimization of the performance of the given cleavage agent. For example, when using cyanogen bromide as a cleavage agent, a highly acidic environment (e.g., pH 2, formic acid) is necessary; however, when using trypsin as a cleavage agent, a slightly alkaline environment (pH 8) is optimal. As a general rule, the pH of the reaction milieu should not alter the chemical integrity of the protein during the digestion and should not change during the course of the fragmentation reaction.
Temperature— A temperature between 25 and 37 is adequate for most digestions. The temperature used is intended to minimize chemical side reactions. The type of protein under test will dictate the temperature of the reaction milieu, because some proteins are more susceptible to denaturation as the temperature of the reaction increases. For example, digestion of recombinant bovine somatropin is conducted at 4, because at higher temperatures it will precipitate during digestion.
Time— If a sufficient amount of sample is available, a time course study is considered in order to determine the optimum time to obtain a reproducible map and avoid incomplete digestion. Time of digestion varies from 2 to 30 hours. The reaction is stopped by the addition of an acid that does not interfere in the tryptic map, or by freezing.
Amount of Cleavage Agent— Although excessive amounts of cleavage agent are used to accomplish a reasonably rapid digestion time (i.e., 6 to 20 hours), the amount of cleavage agent is minimized to avoid its contribution to the chromatographic map pattern. A protein to protease ratio between 20:1 and 200:1 is generally used. It is recommended that the cleavage agent can be added in two or more stages to optimize cleavage. Nonetheless, the final reaction volume remains small enough to facilitate the next step in peptide mapping—the separation step. To sort out digestion artifacts that might be interfering with the subsequent analysis, a blank determination is performed, using a digestion control with all the reagents, except the test protein.
CHROMATOGRAPHIC SEPARATION
Many techniques are used to separate peptides for mapping. The selection of a technique depends on the protein being mapped. Techniques that have been successfully used for separation of peptides are shown in Table 2.
Table 2. Techniques Used for the Separation of Peptides
Reverse-Phase High-Performance Liquid
Chromatography (RP-HPLC)
Ion-Exchange Chromatography (IEC)
Hydrophobic Interaction Chromatography (HIC)
Polyacrylamide Gel Electrophoresis (PAGE), nondenaturating
Sodium Dodecyl Sulfate Polyacrylamide Gel
Electrophoresis (SDS-PAGE)
Capillary Electrophoresis (CE)
Paper Chromatography
High-Voltage Paper Electrophoresis (HVPE)
In this section, a most widely used reverse-phase HPLC (RP-HPLC) method is described as one of the procedures of chromatographic separation.
The purity of solvents and mobile phases is a critical factor in HPLC separation. HPLC-grade solvents and water that are commercially available are recommended for RP-HPLC. Dissolved gases present a problem in gradient systems where the solubility of the gas in a solvent may be less in a mixture than in a single solvent. Vacuum degassing and agitation by sonication are often used as useful degassing procedures. The solid particles in the solvents are drawn into the HPLC system, they can damage the sealing of pump valves or clog the top of the chromatographic column. Both pre- and post-pump filtration is also recommended.
Chromatographic Column— The selection of a chromatographic column is empirically determined for each protein. Columns with 100 or 300 pore size with silica support can give optimal separation. For smaller peptides, octylsilane chemically bonded to totally porous silica articles, 3 to 10 µm in diameter (L7) and octadecylsilane chemically bonded to porous silica or ceramic microparticles, 3 to 10 µm in diameter (L1) column packings are more efficient than the butyl silane chemically bonded to totally porous silica particles, 5 to 10 µm in diameter (L26) packing.
Solvent— The most commonly used solvent is water with acetonitrile as the organic modifier to which less than 0.1% of trifluoroacetic acid is added. If necessary, add isopropyl alcohol or n-propyl alcohol to solubilize the digest components, provided that the addition does not unduly increase the viscosity of the components.
Mobile Phase— Buffered mobile phases containing phosphate are used to provide some flexibility in the selection of pH conditions, since shifts of pH in the 3.0 to 5.0 range enhance the separation of peptides containing acidic residues (e.g., glutamic and aspartic acids). Sodium or potassium phosphates, ammonium acetate, phosphoric acid, and a pH between 2 and 7 (or higher for polymer-based supports) have also been used with acetonitrile gradients. Acetonitrile-containing trifluoroacetic acid is also used quite often.
Gradient Selection— Gradients can be linear, nonlinear, or include step functions. A shallow gradient is recommended in order to separate complex mixtures. Gradients are optimized to provide clear resolution of one or two peaks that will become “marker” peaks for the test.
Isocratic Selection— Isocratic HPLC systems using a single mobile phase are used on the basis of their convenience of use and improved detector responses. Optimal composition of a mobile phase to obtain clear resolution of each peak is sometimes difficult to establish. Mobile phases for which slight changes in component ratios or in pH significantly affect retention times of peaks in peptide maps should not be used in isocratic HPLC systems.
Other Parameters— Temperature control of the column is usually necessary to achieve good reproducibility. The flow rates for the mobile phases range from 0.1 to 2.0 mL per minute, and the detection of peptides is performed with a UV detector at 200 to 230 nm. Other methods of detection have been used (e.g., postcolumn derivatization), but they are not as robust or as versatile as UV detection.
System Suitability— The section System Suitability under Chromatography 621 provides an experimental means for measuring the overall performance of the test method. The acceptance criteria for system suitability depend on the identification of critical test parameters that affect data interpretation and acceptance. These critical parameters are also criteria that monitor peptide digestion and peptide analysis. An indicator that the desired digestion endpoint was achieved is the comparison with a reference standard or reference material, which is treated exactly as the article under test. The use of a USP Reference Standard in parallel with the protein under test is critical in the development and establishment of system suitability limits. In addition, a specimen chromatogram should be included with the USP Reference Standard or reference material for comparison purposes. Other indicators may include visual inspection of protein or peptide solubility, the absence of intact protein, or measurement of responses of a digestion-dependent peptide. The critical system suitability parameters for peptide analysis will depend on the particular mode of peptide separation and detection, and on the data analysis requirements.
When peptide mapping is used as an identification test, the system suitability requirements for the identified peptides covers selectivity and precision. In this case, as well as when identification of variant proteins is done, the identification of the primary structure of the peptide fragments in the peptide map provides both a verification of the known primary structure and the identification of protein variants by comparison with the peptide map of the USP Reference Standard or reference material for the specified protein. The use of a digested USP Reference Standard or reference material for a given protein in the determination of peptide resolution is the method of choice. For an analysis of a variant protein, a characterized mixture of a variant and a reference standard can be used, especially if the variant peptide is located in a less-resolved region of the map. The index of pattern consistency can be simply the number of major peptides detected. Peptide pattern consistency can be best defined by the resolution of peptide peaks. Chromatographic parameters—such as peak-to-peak resolution, maximum peak width, peak tailing factors, and column efficiency—may be used to define peptide resolution. Depending on the protein under test and the method of separation used, single peptide or multiple peptide resolution requirements may be necessary.
The replicate analysis of the digest of the USP Reference Standard or reference material for the protein under test yields measures of precision and quantitative recovery. Recovery of the identified peptides is generally ascertained by the use of internal or external peptide standards. The precision is expressed as the relative standard deviation (RSD). Differences in the recovery and precision of the identified peptides are expected; therefore, the system suitability limits will have to be established for both the recovery and the precision of the identified peptides. These limits are unique for a given protein and will be specified in the individual monograph.
Visual comparison of the relative retention times, the peak responses, the number of peaks, and the overall elution pattern is completed initially. It is then complemented and supported by mathematical analysis of the peak response ratios and by the chromatographic profile of a 1:1 (v/v) mixture of sample and USP Reference Standard or reference material digest. If all peaks in the sample digest and in the USP Reference Standard or reference material digest have the same relative retention times and peak response ratios, then the identity of the sample under test is confirmed.
If peaks that initially eluted with significantly different relative retention times are then observed as single peaks in the 1:1 mixture, the initial difference would be an indication of system variability. However, if separate peaks are observed in the 1:1 mixture, this would be evidence of the nonequivalence of the peptides in each peak. If a peak in the 1:1 mixture is significantly broader than the corresponding peak in the sample and USP Reference Standard or reference material digest, it may indicate the presence of different peptides. The use of computer-aided pattern recognition software for the analysis of peptide mapping data has been proposed and applied, but issues related to the validation of the computer software preclude its use in a compendial test in the near future. Other automated approaches have been used that employ mathematical formulas, models, and pattern recognition. Such approaches, for example, the automated identification of compounds by IR spectroscopy and the application of diode-array UV spectral analysis for identification of peptides, have been proposed. These methods have limitations due to inadequate resolutions, co-elution of fragments, or absolute peak response differences between USP Reference Standard or reference material and sample fragments.
The numerical comparison of the retention times and peak areas or peak heights can be done for a selected group of relevant peaks that have been correctly identified in the peptide maps. Peak areas can be calculated using one peak showing relatively small variation as an internal reference, keeping in mind that peak area integration is sensitive to baseline variation and likely to introduce error in the analysis. Alternatively, the percentage of each peptide peak height relative to the sum of all peak heights can be calculated for the sample under test. The percentage is then compared to that of the corresponding peak of the USP Reference Standard or reference material. The possibility of autohydrolysis of trypsin is monitored by producing a blank peptide map that is the peptide map obtained when a blank solution is treated with trypsin.
The minimum requirement for the qualification of peptide mapping is an approved test procedure that includes system suitability as a test control. In general, for an IND, qualification of peptide mapping for a protein is sufficient. As the regulatory approval process for the protein progresses, additional qualifications of the test can include a partial validation of the analytical procedure to provide assurance that the method will perform as intended in the development of a peptide map for the specified protein.
ANALYSIS and IDENTIFICATION OF PEPTIDES
This section gives guidance on the use of peptide mapping during development in support of regulatory applications.
The use of a peptide map as a qualitative tool does not require the complete characterization of the individual peptide peaks. However, validation of peptide mapping in support of regulatory applications requires rigorous characterization of each of the individual peaks in the peptide map. Methods to characterize peaks range from N-terminal sequencing of each peak followed by amino acid analysis to the use of mass spectroscopy (MS).
For characterization purposes, when N-terminal sequencing and amino acids analysis are used, the analytical separation is scaled up. Since scale-up might affect the resolution of peptide peaks, it is necessary, using empirical data, to assure that there is no loss of resolution due to scale-up. Eluates corresponding to specific peptide peaks are collected, vacuum-concentrated, and chromatographed again, if necessary. Amino acid analysis of fragments may be limited by the peptide size. If the N-terminus is blocked, it may need to be cleared before sequencing. C-terminal sequencing of proteins in combination with carboxypeptidase and MALDITOF-MS can also be used for characterization purposes.
The use of MS for characterization of peptide fragments is by direct infusion of isolated peptides or by the use of on-line LC-MS for structure analysis. In general, it includes electrospray and matrix-assisted laser desorption ionization coupled to time-of-flight analyzer (MALDITOF) as well as fast atom bombardment (FAB). Tandem MS has also been used to sequence a modified protein and to determine the type of amino acid modification that has occurred. The comparison of mass spectra of the digests before and after reduction provides a method to assign the disulfide bonds to the various sulfhydryl-containing peptides.
If regions of the primary structure are not clearly demonstrated by the peptide map, it might be necessary to develop a secondary peptide map. The goal of a validated method of characterization of a protein through peptide mapping is to reconcile and account for at least 95% of the theoretical composition of the protein structure.
The Use of Peptide Mapping for Genetic Stability Evaluation
A validated peptide map can be used to assess the integrity of the predicted primary sequence of a protein product (i.e., its genetic stability). It can also be used to determine lot-to-lot consistency of the biotechnology-derived product process. Furthermore, the performance of the protein expression of the production system is best assessed by peptide mapping of the expressed protein. Peptide maps of protein produced at various times of the protein expression process, including a point well beyond the normal protein expression time, compared with those of a USP Reference Standard or reference material, will evaluate the genetic stability of the expression system as a function of time.
Variant protein sequences can arise from a genetic variation at the DNA level (point mutation) or as an error in the translation process. A validated peptide map is the best approach to the detection of protein variants. However, the limitations of the peptide mapping itself must be taken into consideration. The detection of a structured variant is possible only if the corresponding peptide variant is easily isolated and characterized. To establish genetic stability will require the use of a battery of biochemical methods, provided that the variants have properties different from those of the “normal” protein.
Validation
CRITICAL FACTORS
Validation of peptide mapping requires that a protocol be designed, outlining in detail the experiment to be conducted and the criteria for acceptance of the map. Criteria for acceptance of mapping include detection limit, specificity, linearity, range, accuracy, precision, and reagent stability. Reproducibility of the peptide map is a critical element in the utilization of such a map as an identity test and for confirming genetic stability. Those technical aspects of peptide mapping that influence the reproducibility of the map will be discussed.
The setting of limits, with respect to quantification (peak area or height) and identification (retention times) for the selected group of relevant peaks is based on empirical observations. These limits detect significant differences between the sample and USP Reference Standard or reference material within a series of analyses.
Another critical issue is the recovery of peptides and its impact on peak area determination and reproducibility and on the establishment of acceptance criteria. The recovery criteria address all aspects of test methodology, from digestion to chromatographic conditions. Determination of peptide recovery includes quantitative amino acid analysis, spike addition, radiolabeling, and UV summation. An overall recovery of about 80% is considered satisfactory. Recovery of individual peptides is more problematic and is handled on a case-by-case basis. The critical factors considered in the validation of a peptide map are as follows.
Written Test Procedures— These procedures include a detailed description of the analytical method in which reagents, equipment, sample preparation, method of analysis, and analysis of the data are defined.
Validation Protocol— A protocol is prepared that contains a procedure for test validation.
Acceptance Criteria— The criteria can be minimal at the early stages, but need to be better defined as validation studies progress.
Reporting of Results— Results from the validation study are documented with respect to the analytical parameters listed in the validation protocol.
Revalidation of the Test Procedure— If the method used requires alteration that could affect the analytical parameter previously assessed in the validation of the procedure, the test procedure must be revalidated. Significant changes in the processing of the article, in laboratories performing the analysis, in formulation of the bulk or the finished products, and in any other significant parameter will require revalidation of the methods.
REQUIREMENTS
Precision—
Intratest Precision— This is a measure of the reproducibility of peptide mapping. The two critical steps in peptide mapping are fragmentation (i.e., digestion) and separation of peptides. An acceptable precision occurs where the absolute retention times and the relative peak areas are constant from run to run, and the average variation in retention time is small relative to that of a selected internal reference peak. The reproducibility of the map can be enhanced if a temperature-controlled column oven is used, if an extensive equilibration of the system is performed prior to the start of the test, if a blank (control digest mixture without protein) is run first to minimize “first run effects,” and if a USP Reference Standard or a reference material digest is interspersed periodically with test samples to evaluate chromatographic drift.
The criteria for validation of the fragmentation step are similar to those described below for separation of peptides, but they are met for consecutive tests of a series of separately prepared digests of the protein under test.
The criteria for validation of the separation of peptides step include the following:
  1. The average standard deviation of the absolute retention times of all major peaks for a set of consecutive tests of the same digest does not exceed a specified acceptance criterion.
  2. The average standard deviation of absolute peak area for all fully resolved major peaks does not exceed a specified percentage.
Intertest Precision— This is a measure of the reproducibility of the peptide mapping when the test is performed on different days, by different analysts, in different laboratories, with reagents or enzymes from different suppliers or different lots from the same supplier, with different instruments, on columns of different makes or columns of the same make from different lots, and on individual columns of the same make from the same lot. Although it would be desirable, from a scientific perspective, to validate all of these variables in terms of their impacts on precision, a practical approach is to validate the test using those variables most likely to be encountered under operational conditions. Additional variables can be included when needed.
The experimental design allows the analyst to make comparisons using peak retention times and areas that are expressed relative to a highly reproducible internal reference peak within the same chromatogram. The relative peak area is expressed as the ratio of the peak area to that of the internal reference peak. The relative retention time can be expressed as the difference between the absolute retention time and that of the reference peak. The use of relative values eliminates the need to make separate corrections for differences due to injector-to-injector volumes, units of measure for peak areas, column dimensions, and instrument dead volumes. The variability in the retention times and peak areas for the Intertest Precision experiments is expected to be slightly higher than the variability observed for Intratest Precision.
Robustness— Factors such as composition of the Mobile Phase, protease quality or chemical reagent purity, column variation and age, and digest stability are likely to affect the overall performance of the test and its reproducibility. Tolerances for each of the key parameters are evaluated and baseline limits established in case the test is used for routine lot release purposes.
Mobile Phase— The composition of the Mobile Phase is optimized to obtain the maximum resolution of peptides throughout the elution profile. A balance between optimal resolution and overall reproducibility is desired. A lower pH might improve peak separation but might shorten the life of the column, resulting in lack of reproducibility. Peptide maps at a pH above and below the pH of the procedure are compared to the peptide map obtained at the pH of the procedure and checked for significant differences; they are also reviewed with respect to the acceptance criteria established in the validation protocol.
Protease Quality or Chemical Reagent Purity— A sample of the USP Reference Standard or reference material for the protein under test is prepared and digested with different lots of cleavage agent. The chromatograms for each digest are compared in terms of peak areas, shape, and number. The same procedure can be applied to other critical chemicals or pretreatment procedures used during sample preparation, such as reducing and carboxymethylation reagents.
Column Considerations— Column-to-column variability, even within a single lot, can affect the performance of the column in the development of peptide maps. Column size may also lead to significant differences. A USP Reference Standard or reference material of the protein under test is digested and the digest is chromatographed on different lots of column from a single manufacturer. The maps are then evaluated in terms of the overall elution profile, retention times, selectivity resolution, and recovery. To evaluate the overall lifetime of the column in terms of robustness, perform a peptide mapping test on different columns and vary significantly the number of injections (e.g., from 10 injections to 250 injections). The resulting maps can then be compared for significant differences in peak broadening, peak area, and overall resolution. As a column ages, an increase in back pressure might be observed that might affect the peptide maps.
A sensible precaution in the use of peptide mapping columns is to select alternative columns in case the original columns become unavailable or are discontinued. Perform a peptide mapping test using equivalent columns from different manufacturers, and examine the maps. Differences in particle shape and size, pore size and volume, carbon load, and end-capping can lead to significant differences in retention times, elution profile selectivity, resolution, and recovery. Slight modifications in the gradient profile may be required to achieve equivalency of mapping when using columns from different manufacturers. [NOTE—The equivalency between instrumentation used for the validation of the test and for routine quality control testing should be considered. It might be preferable to use the same HPLC system for all applications. Otherwise, equivalency of the systems is determined, which may require some changes in the chromatographic test conditions.]
Digest Stability— The length of time a digest can be kept before it is chromatographed, as well as the conditions under which the digest is stored before chromatography, is assessed. Several aliquots from a single digest are stored at different storage conditions and chromatographed. These maps are then evaluated for significant differences.
Reproducibility— Determination of various parameters indicated above is repeated using the same USP Reference Standard or reference material and test sample in at least two different laboratories by two analysts equipped with similar HPLC systems. The generated peptide maps are evaluated for significant differences.

POLYACRYLAMIDE GEL ELECTROPHORESIS
Polyacrylamide gel electrophoresis (PAGE) is used for the qualitative characterization of proteins in biological preparations, for control of purity, and for quantitative determinations. This procedure is limited to the analysis of proteins with a weight range of 14,000 to 100,000 Da. It is possible to extend the weight range of an electrophoresis gel by various techniques (e.g., gradient gels or particular buffer systems), but such techniques will not be discussed in this chapter. Analytical gel electrophoresis is an appropriate method with which to identify and to assess the homogeneity of proteins in drug substances. These methods are routinely used for the estimation of protein subunit molecular weights and for the determination of the subunit compositions of purified proteins.
Ready-to-use gels and reagents are commercially available and can be used instead of those described in this chapter, provided that they give equivalent results and that they meet the validation requirements.
General Principle of Electrophoresis
Under the influence of an electrical field, charged particles migrate in the direction of the electrode bearing the opposite polarity. In gel electrophoresis, the movements of the particles are retarded by interactions with the surrounding gel matrix, which acts as a molecular sieve. The opposing interactions of the electrical force and molecular sieving result in differential migration rates according to sizes, shapes, and charges of particles. Because of their different physicochemical properties, different macromolecules of a mixture will migrate at different speeds during electrophoresis and thus will be separated into discrete fractions. Electrophoretic separations can be conducted in systems without support phases (e.g., free solution separation in capillary electrophoresis) and in stabilizing media, such as thin-layer plates, films, or gels.
Characteristics of Polyacrylamide Gels for Protein Electrophoresis
The sieving properties of polyacrylamide gels are established by the three-dimensional network of fibers and pores that is formed as the bifunctional bisacrylamide cross-links adjacent to polyacrylamide chains. Polymerization is catalyzed by a free radical-generating system composed of ammonium persulfate and N,N,N¢,N¢-tetramethylethylenediamine (TEMED).
As the acrylamide concentration of a gel increases, its effective pore size decreases. The effective pore size of a gel is operationally defined by its sieving properties, that is, by the resistance it imparts to the migration of macromolecules. There are limits to the acrylamide concentrations that can be used. At high acrylamide concentrations, gels break much more easily and are difficult to handle. As the pore size of a gel decreases, the migration rate of a protein through the gel decreases. By adjusting the pore size of a gel, through manipulating the acrylamide concentration, the resolution of the method can be optimized for a given protein product. Thus, a given gel is physically characterized by its respective composition of acrylamide and bisacrylamide.
In addition to the composition of the gel, the state of the protein is an important component to the electrophoretic mobility. In the case of proteins, the electrophoretic mobility is dependent on the pK value of the charged groups and the size of the molecule. It is influenced by the type, the concentration, and pH of the buffer, by the temperature and the field strength, and by the nature of the support material.
Denaturation with Sodium Dodecyl Sulfate
Denaturing PAGE using sodium dodecyl sulfate (SDS) is the most common mode of electrophoresis used in assessing the pharmaceutical quality of protein products. Typically, analytical electrophoresis of proteins is carried out under conditions that ensure dissociation of the proteins into their individual polypeptide subunits and that minimize aggregation of these subunits. The strongly anionic detergent SDS is used in combination with heat to dissociate the proteins before they are loaded on the gel. The denatured polypeptides bind SDS, become negatively charged, and exhibit a consistent charge-to-weight ratio regardless of protein type. Because the amount of SDS bound is almost always proportional to the molecular weight of the polypeptide and is typically independent of its sequence, SDS-polypeptide complexes migrate through polyacrylamide gels in reasonable accordance with the size of the polypeptide.
The electrophoretic mobilities of the resultant detergent-polypeptide complexes all assume the same functional relationship to the molecular weights of polypeptides. Migration of SDS derivatives is toward the anode in a predictable manner, with low molecular weight complexes migrating faster than larger ones. This means that the molecular weight of a protein can be estimated from its relative mobility in calibrated SDS-PAGE and that a single band in such a gel is a criterion of purity.
Modifications to the polypeptide backbone, such as N- or O-linked glycosylation, however, have a significant impact on the apparent molecular weight of a protein. This is due to SDS not binding to a carbohydrate moiety in a manner similar to the polypeptide. Thus, a consistent charge-to-weight ratio is not maintained. The apparent molecular weight of proteins having undergone post-translational modifications is not a true reflection of the weight of the polypeptide chain.
REDUCING CONDITIONS
Polypeptide subunits and their three-dimensional structure can be maintained in proteins by the presence of disulfide bonds. A goal of SDS-PAGE analysis under reducing conditions is to disrupt this structure by reducing disulfide bonds. Complete denaturation and dissociation of proteins by treatment with 2-mercaptoethanol or dithiothreitol (DTT) will result in unfolding of the polypeptide backbone and subsequent complexation with SDS. Under these conditions, the molecular weight of the polypeptide subunits can be calculated by linear regression in the presence of suitable molecular weight standards.
NONREDUCING CONDITIONS
For some analyses, complete dissociation of protein to peptide subunits is not desirable. In the absence of treatment with reducing agents, disulfide covalent bonds remain intact, preserving the oligomeric form of the protein. Oligomeric SDS-protein complexes migrate more slowly than their SDS-polypeptide subunits. In addition, nonreduced proteins may not be completely saturated with SDS and, hence, may not bind the detergent in a constant weight ratio. This makes molecular weight determinations of these molecules less straightforward than analyses of fully denatured polypeptides, because it is necessary that both standards and unknown proteins be in similar configurations for valid comparisons. However, the staining of a single band in such a gel is a criterion of purity.
CHARACTERISTICS OF A DISCONTINUOUS BUFFER SYSTEM
The most popular electrophoretic method for the characterization of a complex mixture of proteins involves the use of a discontinuous buffer system consisting of two contiguous, but distinct, gels: a resolving or separating (lower) gel and a stacking (upper) gel. The two gels are cast with different porosities, pHs, and ionic strengths. In addition, different mobile ions are used in the gel and electrode buffers. The buffer discontinuity concentrates large volumes of sample in the stacking gel, resulting in improved resolution. When power is applied, a voltage drop develops across the sample solution that drives the proteins into the stacking gel. Glycinate ions from the electrode buffer follow the proteins into the stacking gel. A moving boundary region is rapidly formed with the highly mobile chloride ions in the front and the relatively slow glycinate ions in the rear. A localized high-voltage gradient forms between the leading and trailing ion fronts, causing the SDS-protein complexes to form into a thin zone (stack) and migrate between the chloride and glycinate phases. Within a broad limit, regardless of the height of the applied sample, all SDS-proteins condense into a very narrow region and enter the resolving gel as a well-defined, thin zone of high protein density. The large-pore stacking gel does not retard the migration of most proteins and serves mainly as an anticonvective medium. At the interface between the stacking and resolving gels, the proteins experience a sharp retardation due to the restrictive pore size of the resolving gel. Once in the resolving gel, proteins continue to be slowed by the sieving of the matrix. The glycinate ions overtake the proteins, which then move in a space of uniform pH formed by the TRIS and glycine. Molecular sieving causes the SDS-polypeptide complexes to separate on the basis of their molecular weights.
PREPARATION OF GELS
In a discontinuous buffer SDS-polyacrylamide gel, it is important to pour the resolving gel, let the gel set, and then pour the stacking gel because the composition of acrylamide-bisacrylamide, buffer, and pH are different.
Gel Stock Solutions—
30% Acrylamide-Bisacrylamide Solution— Prepare a solution containing 290 g of acrylamide and 10 g of methylene bisacrylamide per L of warm water, and filter. [NOTE—Acrylamide and methylene bisacrylamide are slowly converted during storage to acrylic acid and bisacrylic acid, respectively. This deamidation reaction is catalyzed by light and alkali. The pH of the solution must be 7.0 or lower. Store the solution in dark bottles at room temperature. Fresh solutions are prepared every month.]
Ammonium Persulfate Solution— Prepare a small quantity of solution having a concentration of 100 g of ammonium persulfate per L, and store at 4. [NOTE—Ammonium persulfate provides the free radicals that drive polymerization of acrylamide and bisacrylamide. Ammonium persulfate decomposes slowly; therefore, prepare fresh solutions weekly.]
TEMED— Use an electrophoresis-grade reagent. [NOTE— TEMED accelerates the polymerization of acrylamide and bisacrylamide by catalyzing the formation of free radicals from ammonium persulfate. Because TEMED works only as a free base, polymerization is inhibited at low pH.]
SDS Solution— Use an electrophoresis-grade reagent. Prepare a solution having a concentration of about 100 g of SDS per L, and store at room temperature.
1.5 M Buffer Solution— Transfer about 90.8 g of tris(hydroxymethyl)aminomethane (TRIS) to a 500-mL flask, dissolve in 400 mL of water, adjust with hydrochloric acid to a pH of 8.8, dilute with water to volume, and mix.
1 M Buffer Solution— Transfer about 60.6 g of TRIS to a 500-mL flask, add 400 mL of water, adjust with hydrochloric acid to a pH of 6.8, dilute with water to volume, and mix.
Plate Preparation— Clean two glass plates (10 cm × 8 cm), the polytef comb, the two spacers, and the silicone rubber tubing (0.6 mm × 35 cm) with mild detergent, rinse thoroughly with water, and blot dry.
Lubricate the spacers and the tubing with nonsilicone grease. Apply the spacers along each of the two short sides of the glass plate 2 mm away from the edges and 2 mm away from the long side corresponding to the bottom of the gel.
Begin to lay the tubing on the glass plate by using one spacer as a guide. Carefully twist the tubing at the bottom of the spacer, and follow the long side of the glass plate. While holding the tubing with one finger along the long side, twist the tubing again, and lay it on the second short side of the glass plate, using the spacer as a guide.
Place the second glass plate in perfect alignment, and hold the mold together by hand pressure. Apply two clamps on each of the two short sides of the mold. Carefully apply four clamps on the longer side of the gel mold, thus forming the bottom of the gel mold. Verify that the tubing is running along the edge of the glass plates and has not been extruded while placing the clamps. The gel mold is now ready for pouring the gel.
Resolving Gel— In a conical flask, prepare the appropriate volume of solution containing the desired concentration of acrylamide as directed in Table 3. Mix the components in the order shown. Before adding the Ammonium Persulfate Solution and the TEMED, pour the solution in a disposable filtration unit equipped with a nitrocellulose filter having a 0.45-µm porosity, and apply vacuum. Allow the solution to degas by swirling the filtration unit, and disconnect the vacuum when no more bubbles are formed in the solution. Add appropriate amounts of Ammonium Persulfate Solution and TEMED, as directed in Table 3, swirl, and pour immediately into the gap between the two glass plates of the mold. Leave sufficient space for the stacking gel (the length of the teeth of the comb plus 1 cm). Using a pipet, carefully overlay the solution with water-saturated isobutyl alcohol. Leave the gel in a vertical position at room temperature for polymerization.
Table 3. Preparation of Resolving Gel
Component Volumes (mL) per Gel Mold Volume Below
Solution Components 5 mL l0 mL 15 mL 20 mL 25 mL 30 mL 40 mL 50 mL
6% Acrylamide
Water 2.6 5.3 7.9 10.6 13.2 15.9 21.2 26.5
30% Acrylamide-Bisacrylamide Solution 1.0 2.0 3.0 4.0 5.0 6.0 8.0 10.0
1.5 M Buffer Solution 1.3 2.5 3.8 5.0 6.3 7.5 10.0 12.5
SDS Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
Ammonium Persulfate Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
TEMED 0.004 0.008 0.012 0.016 0.02 0.024 0.032 0.04
8% Acrylamide
Water 2.3 4.6 6.9 9.3 11.5 13.9 18.5 23.2
30% Acrylamide-Bisacrylamide Solution 1.3 2.7 4.0 5.3 6.7 8.0 10.7 13.3
1.5 M Buffer Solution 1.3 2.5 3.8 5.0 6.3 7.5 10.0 12.5
SDS Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
Ammonium Persulfate Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
TEMED 0.003 0.006 0.009 0.012 0.015 0.018 0.024 0.03
10% Acrylamide
Water 1.9 4.0 5.9 7.9 9.9 11.9 15.9 19.8
30% Acrylamide-Bisacrylamide Solution 1.7 3.3 5.0 6.7 8.3 10.0 13.3 16.7
1.5 M Buffer Solution 1.3 2.5 3.8 5.0 6.3 7.5 10.0 12.5
SDS Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
Ammonium Persulfate Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
TEMED 0.002 0.004 0.006 0.008 0.01 0.012 0.016 0.02
12% Acrylamide
Water 1.6 3.3 4.9 6.6 8.2 9.9 13.2 16.5
30% Acrylamide-Bisacrylamide Solution 2.0 4.0 6.0 8.0 10.0 12.0 16.0 20.0
1.5 M Buffer Solution 1.3 2.5 3.8 5.0 6.3 7.5 10.0 12.5
SDS Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
Ammonium Persulfate Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
TEMED 0.002 0.004 0.006 0.008 0.01 0.012 0.016 0.02
14% Acrylamide
Water 1.4 2.7 3.9 5.3 6.6 8.0 10.6 13.8
30% Acrylamide-Bisacrylamide Solution 2.3 4.6 7.0 9.3 11.6 13.9 18.6 23.2
1.5 M Buffer Solution 1.2 2.5 3.6 5.0 6.3 7.5 10.0 12.5
SDS Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
Ammonium Persulfate Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
TEMED 0.002 0.004 0.006 0.008 0.01 0.012 0.016 0.02
15% Acrylamide
Water 1.1 2.3 3.4 4.6 5.7 6.9 9.2 11.5
30% Acrylamide-Bisacrylamide Solution 2.5 5.0 7.5 10.0 12.5 15.0 20.0 25.0
1.5 M Buffer Solution 1.3 2.5 3.8 5.0 6.3 7.5 10.0 12.5
SDS Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
Ammonium Persulfate Solution 0.05 0.1 0.15 0.2 0.25 0.3 0.4 0.5
TEMED 0.002 0.004 0.006 0.008 0.01 0.012 0.016 0.02
After polymerization is complete (about 30 minutes later), pour off the overlay, and wash the top of the gel several times with water to remove the isobutyl alcohol overlay and any unpolymerized acrylamide. Drain as much fluid as possible from the top of the gel, and then remove any remaining water with the edge of a paper towel.
Stacking Gel— In a conical flask, prepare the appropriate volume of solution containing the desired concentration of acrylamide, as directed in Table 4. Mix the components in the order shown. Before adding the Ammonium Persulfate Solution and the TEMED, pour the solution into a disposable filtration unit equipped with a nitrocellulose filter having a 0.45-µm porosity, and apply vacuum. Allow the solution to degas by swirling the filtration unit, and disconnect the vacuum when no more bubbles are formed in the solution. Add appropriate amounts of Ammonium Persulfate Solution and TEMED as directed in Table 4, swirl, and pour immediately into the gap between the two glass plates of the mold directly onto the surface of the polymerized Resolving Gel. Immediately insert a clean polytef comb into the stacking gel solution, being careful to avoid trapping air bubbles. Add more stacking gel solution to fill the spaces of the comb completely. Leave the gel in a vertical position, and allow to polymerize at room temperature. After polymerization is complete (about 30 minutes later), carefully remove the polytef comb, and proceed as directed below.
Table 4. Preparation of Stacking Gel
Component Volumes (mL) per Gel Mold Volume Below
Solution Components 1 mL 2 mL 3 mL 4 mL 5 mL 6 mL 8 mL 10 mL
Water 0.68 1.4 2.1 2.7 3.4 4.1 5.5 6.8
30% Acrylamide-Bisacrylamide Solution 0.17 0.33 0.5 0.67 0.83 1.0 1.3 1.7
1.0 M Buffer Solution 0.13 0.25 0.38 0.5 0.63 0.75 1.0 1.25
SDS Solution 0.01 0.02 0.03 0.04 0.05 0.06 0.08 0.1
Ammonium Persulfate Solution 0.01 0.02 0.03 0.04 0.05 0.06 0.08 0.1
TEMED 0.001 0.002 0.003 0.004 0.005 0.006 0.008 0.01
ELECTROPHORETIC SEPARATION
Sample Buffer 1— Dissolve 1.89 g of TRIS, 5.0 g of SDS, 50 mg bromophenol blue, and 25.0 mL glycerol in 100 mL of water. Adjust with hydrochloric acid to a pH of 6.8, and dilute with water to 125 mL. Before use, dilute with an equal volume of water or sample, and mix.
Sample Buffer 2 (for reducing conditions) Prepare as directed under Sample Buffer 1 except to add 12.5 mL of 2-mercaptoethanol before adjusting the pH. Alternatively, prepare as directed for Sample Buffer 1 except to start with about 1.93 g of TRIS and add a suitable quantity of DTT to obtain a final 100 µM DTT concentration.
Running Buffer— Dissolve 151.4 g of TRIS, 721.0 g of aminoacetic acid, and 50.0 g of SDS in water, dilute with water to 5000 mL, and mix to obtain a stock solution. Immediately before use, dilute this stock solution with water to 10 times its volume, mix, and adjust to a pH between 8.1 and 8.8.
Procedure— Rinse the wells immediately with water or with the Running Buffer to remove any unpolymerized acrylamide. (If necessary, straighten the teeth of the Stacking Gel with a blunt hypodermic needle attached to a syringe.) Remove the clamps on one short side, carefully pull out the tubing, and replace the clamps. Proceed similarly on the other short side. Remove the tubing from the bottom part of the gel.
Mount the completed gel in the electrophoresis apparatus. Add the electrophoresis buffers to the top and bottom reservoirs. Remove any bubbles that become trapped at the bottom of the gel between the glass plates. [NOTES—This is best done with a bent hypodermic needle attached to a syringe. Never pre-run the gel before loading the samples, because this will destroy the discontinuity of the buffer systems. Before loading the sample, carefully rinse the slot with Running Buffer.]
Prepare the test and standard solutions in the recommended Sample Buffer, and treat as directed in the individual monograph. Apply the appropriate volume of each solution to the Stacking Gel wells.
Start the electrophoresis using the conditions recommended by the manufacturer of the equipment. Electrophoresis running time and current or voltage may need to be varied in order to achieve optimum separation. Check that the dye front is moving into the resolving gel. When the dye has reached the bottom of the gel, stop the electrophoresis. Remove the gel assembly from the apparatus, and separate the glass plates. Remove the spacers, cut off and discard the Stacking Gel, and immediately proceed with staining.
DETECTION OF PROTEINS IN GELS
Coomassie staining is the most common protein staining method with a detection level in the order of 1 to 10 µg of protein per band. Silver staining is the most sensitive method for staining proteins in gels, as a band containing 10 to 100 ng can be detected, but the method is more cumbersome and less rugged. All of the steps in gel staining are performed at room temperature with gentle agitation (e.g., on a rocking platform shaker, or equivalent). Gloves must be worn when staining the gels to prevent fingerprint residue staining.
Reagents—
Coomassie Staining Solution— Prepare a solution of Coomassie brilliant blue R-250 having a concentration of 1.25 g per L in a mixture of water, methanol, and glacial acetic acid (5:4:1). Filter, and store at room temperature.
Destaining Solution— Prepare a mixture of water, methanol, and glacial acetic acid (5:4:1).
Fixing Solution 1— Prepare a mixture of water, methanol, and trichloroacetic acid (5:4:1).
Fixing Solution 2— Transfer 250 mL of methanol to a 500-mL volumetric flask, add 0.27 mL of formaldehyde, dilute with water to volume, and mix.
Silver Nitrate Reagent— To a mixture of 40 mL of 1 M sodium hydroxide and 3 mL of ammonium hydroxide, add, dropwise, 8 mL of a 200 g per L solution of silver nitrate with stirring; dilute to 200 mL with water; and mix.
Developing Solution— Transfer 2.5 mL of a citric acid solution (2 in 100) and 0.27 mL of formaldehyde to a 500.0-mL volumetric flask, dilute with water to volume, and mix.
Stopping Solution— Prepare a 10% (v/v) solution of acetic acid.
Coomassie Staining— Immerse the gel in an excess of Coomassie Staining Solution, and incubate for at least 1 hour. Remove the Coomassie Staining Solution. Destain the gel with an excess of Destaining Solution. Change the Destaining Solution several times, until the stained protein bands are clearly distinguishable on a clear background. The more thoroughly the gel is destained, the smaller the amount of protein that can be detected. Destaining can be accelerated by including a few g of anion-exchange resin or a small sponge in the Destaining Solution. [NOTE—The acid-alcohol solutions used in this procedure do not completely fix proteins in the gel. This can lead to losses of some low molecular weight proteins during the staining and destaining of thin gels. Permanent fixation is obtainable by incubating the gel in Fixing Solution 1 for 1 hour before it is immersed in the Coomassie Staining Solution.]
Silver Staining— Immerse the gel in an excess of Fixing Solution 2, and incubate for 1 hour. Remove Fixing Solution 2, add fresh Fixing Solution 2, and incubate for at least 1 hour, or overnight if convenient. Discard Fixing Solution 2, and wash the gel in an excess of water for 1 hour. Soak the gel for 15 minutes in a 1% solution of glutaraldehyde (v/v). Wash the gel twice, for 15 minutes each time, with an excess of water. Soak the gel in fresh Silver Nitrate Reagent for 15 minutes, in darkness. Wash the gel three times, for 5 minutes each time, with an excess of water. Immerse the gel for about 1 minute in Developing Solution until satisfactory staining has been obtained. Stop the development by incubation in the Stopping Solution for 15 minutes, then rinse the gel with water, and proceed with drying as indicated below.
DRYING OF GELS
For Coomassie staining, after the destaining step, incubate the gel in a glycerol solution (1 in 10) for at least 2 hours. For silver staining, add to the final rinsing step a 5-minute incubation in a glycerol solution (1 in 50).
Immerse two sheets of porous cellophane in water, and incubate for 5 to 10 minutes. Place one of the sheets on a drying frame. Carefully lift the gel and place it on the cellophane sheet. Remove any trapped air bubbles, and pour a few mL of water around the edges of the gel. Place the second sheet on top, and remove any trapped air bubbles. Complete the assembly of the drying frame. Place in a drying oven, leave at room temperature until dry, or use a commercial gel dryer.
MOLECULAR WEIGHT DETERMINATION
Molecular weights of proteins are determined by comparison of their mobilities with those of several marker proteins of known molecular weight. Mixtures of proteins with precisely known molecular weights blended for uniform staining are available for calibrating gels. They are available in various molecular weight ranges. Concentrated stock solutions of proteins of known molecular weight are diluted in a sample buffer and loaded on the same gel as the protein sample to be tested.
Immediately after the gel has been run, the position of the bromophenol blue tracking dye is marked to identify the leading edge of the electrophoretic ion front. This can be done by cutting notches in the edges of the gel or by inserting a needle soaked in india ink into the gel at the dye front. After staining, measure the migration distances of each protein band (markers and unknowns) from the top of the Resolving Gel. Divide the migration distance of each protein by the distance traveled by the tracking dye. The normalized migration distances so obtained are called the relative mobilities of the proteins (relative to the dye front) and conventionally denoted as RF. Construct a (semilogarithmic) plot of the logarithm of the molecular weights (MR) of the protein standards as functions of the RF values. [NOTE—The graphs are slightly sigmoid.] From the graph so obtained, estimate the unknown molecular weights by linear regression analysis or interpolation, as long as unknown samples are positioned along the linear part of the graph.
If the proteins of the molecular weight marker are not distributed along 80% of the length of the gel and over the required separation range (i.e., the range covering the product and its dimer or the products and its related impurities), and the separation obtained for the relevant protein bands does not show a linear relationship between the logarithm of the molecular weight and the RF, then the test is not valid.
QUANTIFICATION OF IMPURITIES
Where the impurity limit is specified in the individual monograph, a reference solution corresponding to that level of impurity is prepared by diluting the test solution. For example, where the limit is 5.0%, a reference solution is a 1 in 20 dilution of the test solution. No impurity—any band other than the main band—in the electrophoretogram obtained from the test solution is more intense than the main band obtained with the reference solution.
Under validated conditions and when using the Coomassie staining procedure, impurities may be quantified by normalization to the main band using an integrating densitometer. In this case, the responses must be validated for linearity.

TOTAL PROTEIN ASSAY
The following procedures are provided as illustrations of the determination of total protein content in pharmacopeial preparations. Other techniques, such as HPLC, are also acceptable if total protein recovery is demonstrated. Many of the total protein assay methods described below can be performed successfully using kits from commercial sources. [NOTE—Where water is required, use distilled water.]
Method 1
Protein in solution absorbs UV light at a wavelength of 280 nm, due to the presence of aromatic amino acids, mainly tyrosine and tryptophan. This property is the basis of this method. Protein determination at 280 nm is mainly a function of the tyrosine and tryptophan content of the protein. If the buffer used to dissolve the protein has a high absorbance relative to that of water, there is an interfering substance in the buffer. This interference can be compensated for when the spectrophotometer is adjusted to zero buffer absorbance. If the interference results in a large absorbance that challenges the limit of sensitivity of the spectrophotometer, the results may be compromised. Furthermore, at low concentrations protein can be absorbed onto the cuvette, thereby reducing the content in solution. This can be prevented by preparing samples at higher concentrations or by using a nonionic detergent in the preparation. [NOTE—Keep the Test Solution, the Standard Solution, and the buffer at the same temperature during testing.]
Test Solution— Dissolve a suitable quantity of the protein under test in the appropriate buffer to obtain a solution having a concentration of 0.2 to 2 mg per mL.
Standard Solution— Unless otherwise specified in the individual monograph prepare a solution of USP Reference Standard or reference material for the protein under test in the same buffer and at the same concentration as the Test Solution.
Procedure— Concomitantly determine the absorbances of the Standard Solution and the Test Solution in quartz cells at a wavelength of 280 nm, with a suitable spectrophotometer (see Spectrophotometry and Light-Scattering 851), using the buffer as the blank. To obtain accurate results, the response should be linear in the range of protein concentrations to be assayed.
Light-Scattering— The accuracy of the UV spectroscopic determination of protein can be decreased by the scattering of light by the test specimen. If the proteins in solution exist as particles comparable in size to the wavelength of the measuring light (250 to 300 nm), scattering of the light beam results in an apparent increase in absorbance of the test specimen. To calculate the absorbance at 280 nm due to light-scattering, determine the absorbances of the Test Solution at wavelengths of 320, 325, 330, 335, 340, 345, and 350 nm. Using the linear regression method, plot the log of the observed absorbance versus the log of the wavelength, and determine the standard curve best fitting the plotted points. From the graph so obtained, extrapolate the absorbance value due to light-scattering at 280 nm. Subtract the absorbance from light-scattering from the total absorbance at 280 nm to obtain the absorbance value of the protein in solution. Filtration with a filter having a 0.2-µm porosity or clarification by centrifugation may be performed to reduce the effect of light-scattering, especially if the solution is noticeably turbid.
Calculations— Calculate the concentration, CU, of protein in the test specimen by the formula:
CS(AU / AS),
in which CS is the concentration of the Standard Solution; and AU and AS are the corrected absorbances of the Test Solution and the Standard Solution, respectively (see Spectrophotometry and Light-Scattering 851).
Method 2
This method, commonly referred to as the Lowry assay, is based on the reduction by protein of the phosphomolybdic-tungstic mixed acid chromogen in the Folin-Ciocalteu's phenol reagent, resulting in an absorbance maximum at 750 nm. The Folin-Ciocalteu's phenol reagent reacts primarily with tyrosine residues in the protein, which can lead to variation in the response of the assay to different proteins. Because the method is sensitive to interfering substances, a procedure for precipitation of the protein from the test specimen may be used. Where separation of interfering substances from the protein in the test specimen is necessary, proceed as directed below for Interfering Substances prior to preparation of the Test Solution. The effect of interfering substances can be minimized by dilution, provided the concentration of the protein under test remains sufficient for accurate measurement.
Standard Solutions— Unless otherwise specified in the individual monograph, dissolve the USP Reference Standard or reference material for the protein under test in the buffer used to prepare the Test Solution. Dilute portions of this solution with the same buffer to obtain not fewer than five Standard Solutions having concentrations between 5 and 100 µg of protein per mL, the concentrations being evenly spaced.
Test Solution— Dissolve a suitable quantity of the protein under test in the appropriate buffer to obtain a solution having a concentration within the range of the concentrations of the Standard Solutions. An appropriate buffer will produce a pH in the range of 10.0 to 10.5.
Blank— Use the buffer used for the Test Solution and the Standard Solutions.
Reagents and Solutions—
Copper Sulfate Reagent— Dissolve 100 mg of cupric sulfate and 200 mg of sodium tartrate in water, dilute with water to 50 mL, and mix. Dissolve 10 g of sodium carbonate in water to a final volume of 50 mL, and mix. Slowly pour the sodium carbonate solution into the copper sulfate solution with mixing. Prepare this solution fresh daily.
SDS Solution— Dissolve 5 g of sodium dodecyl sulfate in water, and dilute with water to 100 mL.
Sodium Hydroxide Solution— Dissolve 3.2 g of sodium hydroxide in water, dilute with water to 100 mL, and mix.
Alkaline Copper Reagent— Prepare a mixture of Copper Sulfate Reagent, SDS Solution, and Sodium Hydroxide Solution (1:2:1). This reagent may be stored at room temperature for up to 2 weeks.
Diluted Folin-Ciocalteu’s Phenol Reagent— Mix 10 mL of Folin-Ciocalteu’s phenol TS with 50 mL of water. Store in an amber bottle, at room temperature.
Procedure— To 1 mL of each Standard Solution, the Test Solution, and the Blank, add 1 mL of Alkaline Copper Reagent, and mix. Allow to stand at room temperature for 10 minutes. Add 0.5 mL of the Diluted Folin-Ciocalteu's Phenol Reagent to each solution, mix each tube immediately, and allow to stand at room temperature for 30 minutes. Determine the absorbances of the solutions from the Standard Solutions and the Test Solution at the wavelength of maximum absorbance at 750 nm, with a suitable spectrophotometer (see Spectrophotometry and Light-Scattering 851), using the solution from the Blank to set the instrument to zero.
Calculations— [NOTE—The relationship of absorbance to protein concentration is nonlinear; however, if the standard curve concentration range is sufficiently small, it will approach linearity.] Using the linear regression method, plot the absorbances of the solutions from the Standard Solutions versus the protein concentrations, and determine the standard curve best fitting the plotted points. From the standard curve so obtained and the absorbance of the Test Solution, determine the concentration of protein in the Test Solution.
INTERFERING SUBSTANCES
In the following procedure, deoxycholate-trichloroacetic acid is added to a test specimen to remove interfering substances by precipitation of proteins before testing. This technique also can be used to concentrate proteins from a dilute solution.
Sodium Deoxycholate Reagent— Prepare a solution of sodium deoxycholate in water having a concentration of 150 mg in 100 mL.
Trichloroacetic Acid Reagent— Prepare a solution of trichloroacetic acid in water having a concentration of 72 g in 100 mL.
Procedure— Add 0.1 mL of Sodium Deoxycholate Reagent to 1 mL of a solution of the protein under test. Mix on a vortex mixer, and allow to stand at room temperature for 10 minutes. Add 0.1 mL of Trichloroacetic Acid Reagent, and mix on a vortex mixer. Centrifuge at 3000 × g for 30 minutes, decant the liquid, and remove any residual liquid with a pipet. Redissolve the protein pellet in 1 mL of Alkaline Copper Reagent. Proceed as directed for the Test Solution.
NOTE—Color development reaches a maximum in 20 to 30 minutes during incubation at room temperature, after which there is a gradual loss of color. Most interfering substances cause a lower color yield; however, some detergents cause a slight increase in color. A high salt concentration may cause a precipitate to form. Because different protein species may give different color response intensities, the standard protein and test protein should be the same.
Method 3
This method, commonly referred to as the Bradford assay, is based on the absorption shift from 470 nm to 595 nm observed when the brilliant blue G dye binds to protein. The brilliant blue G dye binds most readily to arginyl and lysyl residues in the protein, which can lead to variation in the response of the assay to different proteins.
Standard Solutions— Unless otherwise specified in the individual monograph, dissolve the USP Reference Standard or reference material for the protein under test in the buffer used to prepare the Test Solution. Dilute portions of this solution with the same buffer to obtain not fewer than five Standard Solutions having concentrations between 100 µg and 1 mg of protein per mL, the concentrations being evenly spaced.
Test Solution— Dissolve a suitable quantity of the protein under test in the appropriate buffer to obtain a solution having a concentration within the range of the concentrations of the Standard Solutions.
Blank— Use the buffer used to prepare the Test Solution and the Standard Solutions.
Coomassie Reagent— Dissolve 100 mg of brilliant blue G3 in 50 mL of alcohol. [NOTE—Not all dyes have the same brilliant blue G content, and different products may give different results.] Add 100 mL of phosphoric acid, dilute with water to 1 L, and mix. Pass the solution through filter paper (Whatman No. 1 or equivalent), and store the filtered reagent in an amber bottle at room temperature. [NOTE—Slow precipitation of the dye will occur during storage of the reagent. Filter the reagent before use.]
Procedure— Add 5 mL of the Coomassie Reagent to 100 µL of each Standard Solution, the Test Solution, and the Blank, and mix by inversion. Avoid foaming, which will lead to poor reproducibility. Determine the absorbances of the solutions from the Standard Solutions and the Test Solution at 595 nm, with a suitable spectrophotometer (see Spectrophotometry and Light-Scattering 851), using the Blank to set the instrument to zero. [NOTE—Do not use quartz (silica) spectrophotometer cells: the dye binds to this material. Because different protein species may give different color response intensities, the standard protein and test protein should be the same.]
There are relatively few interfering substances, but detergents and ampholytes in the test specimen should be avoided. Highly alkaline specimens may interfere with the acidic reagent.
Calculations— [NOTE—The relationship of absorbance to protein concentration is nonlinear; however, if the standard curve concentration range is sufficiently small, it will approach linearity.] Using the linear regression method, plot the absorbances of the solutions from the Standard Solutions versus the protein concentrations, and determine the standard curve best fitting the plotted points. From the standard curve so obtained and the absorbance of the Test Solution, determine the concentration of protein in the Test Solution.
Method 4
This method, commonly referred to as the bicinchoninic acid or BCA assay, is based on reduction of the cupric (Cu2+) ion to cuprous (Cu1+) ion by protein. The bicinchoninic acid reagent is used to detect the cuprous ion. The method has few interfering substances. When interfering substances are present, their effect may be minimized by dilution, provided that the concentration of the protein under test remains sufficient for accurate measurement.
Standard Solutions— Unless otherwise specified in the individual monograph, dissolve the USP Reference Standard or reference material for the protein under test in the buffer used to prepare the Test Solution. Dilute portions of this solution with the same buffer to obtain not fewer than five Standard Solutions having concentrations between 10 and 1200 µg of protein per mL, the concentrations being evenly spaced.
Test Solution— Dissolve a suitable quantity of the protein under test in the appropriate buffer to obtain a solution having a concentration within the range of the concentrations of the Standard Solutions.
Blank— Use the buffer used to prepare the Test Solution and the Standard Solutions.
Reagents—
BCA Reagent— Dissolve about 10 g of bicinchoninic acid, 20 g of sodium carbonate monohydrate, 1.6 g of sodium tartrate, 4 g of sodium hydroxide, and 9.5 g of sodium bicarbonate in water. Adjust, if necessary, with sodium hydroxide or sodium bicarbonate to a pH of 11.25. Dilute with water to 1 L, and mix.
Copper Sulfate Reagent— Dissolve about 2 g of cupric sulfate in water to a final volume of 50 mL.
Copper-BCA Reagent— Mix 1 mL of Copper Sulfate Reagent and 50 mL of BCA Reagent.
Procedure— Mix 0.1 mL of each Standard Solution, the Test Solution, and the Blank with 2 mL of the Copper-BCA Reagent. Incubate the solutions at 37 for 30 minutes, note the time, and allow to come to room temperature. Within 60 minutes following the incubation time, determine the absorbances of the solutions from the Standard Solutions and the Test Solution in quartz cells at 562 nm, with a suitable spectrophotometer (see Spectrophotometry and Light-Scattering 851), using the Blank to set the instrument to zero. After the solutions are cooled to room temperature, the color intensity continues to increase gradually. If substances that will cause interference in the test are present, proceed as directed for Interfering Substances under Method 2. Because different protein species may give different color response intensities, the standard protein and test protein should be the same.
Calculations— [NOTE—The relationship of absorbance to protein concentration is nonlinear; however, if the standard curve concentration range is sufficiently small, it will approach linearity.] Using the linear regression method, plot the absorbances of the solutions from the Standard Solutions versus the protein concentrations, and determine the standard curve best fitting the plotted points. From the standard curve so obtained and the absorbance of the Test Solution, determine the concentration of protein in the Test Solution.
Method 5
This method, commonly referred to as the Biuret assay, is based on the interaction of cupric (Cu2+) ion with protein in an alkaline solution and the resultant development of absorbance at 545 nm.
Standard Solutions— Unless otherwise specified in the individual monograph, prepare a solution of Albumin Human for which the protein content has been previously determined by nitrogen analysis (using the nitrogen-to-protein conversion factor of 6.25) or of the USP Reference Standard or reference material for the protein under test in sodium chloride solution (9 in 1000). Dilute portions of this solution with sodium chloride solution (9 in 1000) to obtain not fewer than three Standard Solutions having concentrations between 0.5 and 10 mg per mL, the concentrations being evenly spaced. [NOTE—Low responses may be observed if the sample under test has significantly different level of proline than that of Albumin Human. A different standard protein may be employed in such cases.]
Test Solution— Prepare a solution of the test protein in sodium chloride solution (9 in 1000) having a concentration within the range of the concentrations of the Standard Solutions.
Blank— Use sodium chloride solution (9 in 1000).
Biuret Reagent— Dissolve about 3.46 g of cupric sulfate in 10 mL of hot water, and allow to cool (Solution 1). Dissolve about 34.6 g of sodium citrate dihydrate and 20.0 g of sodium carbonate in 80 mL of hot water, and allow to cool (Solution 2). Mix Solution 1 and Solution 2, and dilute with water to 200 mL. This Biuret Reagent is stable at room temperature for 6 months. Do not use the reagent if it develops turbidity or contains any precipitate.
Procedure— To one volume of a solution of the Test Solution add an equal volume of sodium hydroxide solution (6 in 100), and mix. Immediately add a volume of Biuret Reagent equivalent to 0.4 volume of the Test Solution, and mix. Allow to stand at a temperature between 15 and 25 for not less than 15 minutes. Within 90 minutes after the addition of the Biuret Reagent, determine the absorbances of the Standard Solutions and the solution from the Test Solution at the wavelength of maximum absorbance at 545 nm, with a suitable spectrophotometer (see Spectrophotometry and Light-Scattering 851), using the Blank to set the instrument to zero. [NOTE—Any solution that develops turbidity or a precipitate is not acceptable for calculation of protein concentration.]
Calculations— Using the least-squares linear regression method, plot the absorbances of the Standard Solutions versus the protein concentrations, determine the standard curve best fitting the plotted points, and calculate the correlation coefficient for the line. [NOTE—Within the given range of the standards, the relationship of absorbance to protein concentration is approximately linear.] A suitable system is one that yields a line having a correlation coefficient of not less than 0.99. From the standard curve and the absorbance of the Test Solution, determine the concentration of protein in the test specimen, making any necessary correction.
Interfering Substances— To minimize the effect of interfering substances, the protein can be precipitated from the initial test specimen as follows. Add 0.1 volume of 50 percent trichloroacetic acid to 1 volume of a solution of the test specimen, withdraw the supernatant layer, and dissolve the precipitate in a small volume of 0.5 N sodium hydroxide. Use the solution so obtained to prepare the Test Solution.
Comments— This test shows minimal difference between equivalent IgG and albumin samples. Addition of the sodium hydroxide and the Biuret Reagent as a combined reagent, insufficient mixing after the addition of the sodium hydroxide, or an extended time between the addition of the sodium hydroxide solution and the addition of the Biuret Reagent will give IgG samples a higher response than albumin samples. The trichloroacetic acid method used to minimize the effects of interfering substances can also be used to determine the protein content in test specimens at concentrations below 500 µg per mL.
Method 6
This fluorometric method is based on the derivatization of the protein with o-phthalaldehyde (OPA), which reacts with the primary amines of the protein (i.e., NH2-terminal amino acid and the -amino group of the lysine residues). The sensitivity of the test can be increased by hydrolyzing the protein before testing. Hydrolysis makes the -amino group of the constituent amino acids of the protein available for reaction with the o-phthalaldehyde reagent. The method requires very small quantities of the protein.
Primary amines, such as tris(hydroxymethyl)aminomethane and amino acid buffers, react with o-phthalaldehyde and must be avoided or removed. Ammonia at high concentrations will react with o-phthalaldehyde as well. The fluorescence obtained when amine reacts with o-phthalaldehyde can be unstable. The use of automated procedures to standardize this procedure may improve the accuracy and precision of the test.
Standard Solutions— Unless otherwise specified in the individual monograph, dissolve the USP Reference Standard or reference material for the protein under test in the buffer used to prepare the Test Solution. Dilute portions of this solution with the same buffer to obtain not fewer than five Standard Solutions having concentrations between 10 and 200 µg of protein per mL, the concentrations being evenly spaced.
Test Solution— Dissolve a suitable quantity of the protein under test in the appropriate buffer to obtain a solution having a concentration within the range of the concentrations of the Standard Solutions.
Blank— Use the buffer used to prepare the Test Solution and the Standard Solutions.
Reagents—
Borate Buffer— Dissolve about 61.83 g of boric acid in water, and adjust with potassium hydroxide to a pH of 10.4. Dilute with water to 1 L, and mix.
Stock OPA Reagent— Dissolve about 120 mg of o-phthalaldehyde in 1.5 mL of methanol, add 100 mL of Borate Buffer, and mix. Add 0.6 mL of polyoxyethylene (23) lauryl ether, and mix. This solution is stable at room temperature for at least 3 weeks.
OPA Reagent— To 5 mL of Stock OPA Reagent add 15 µL of 2-mercaptoethanol. Prepare at least 30 minutes prior to use. This reagent is stable for one day.
Procedure— Adjust each of the Standard Solutions and the Test Solution to a pH between 8 and 10.5. Mix 10 µL of the Test Solution and each of the Standard Solutions with 100 µL of OPA Reagent, and allow to stand at room temperature for 15 minutes. Add 3 mL of 0.5 N sodium hydroxide, and mix. Using a suitable fluorometer (see Spectrophotometry and Light-Scattering 851), determine the fluorescent intensities of solutions from the Standard Solutions and the Test Solution at an excitation wavelength of 340 nm and an emission wavelength between 440 and 455 nm. [NOTE—The fluorescence of an individual specimen is read only once because irradiation decreases the fluorescent intensity.]
Calculations— The relationship of fluorescence to protein concentration is linear. Using the linear regression method, plot the fluorescent intensities of the solutions from the Standard Solutions versus the protein concentrations, and determine the standard curve best fitting the plotted points. From the standard curve so obtained and the fluorescent intensity of the Test Solution, determine the concentration of protein in the test specimen.
Method 7
This method is based on nitrogen analysis as a means of protein determination. Interference caused by the presence of other nitrogen-containing substances in the test specimen can affect the determination of protein by this method. Nitrogen analysis techniques destroy the protein under test but are not limited to protein presentation in an aqueous environment.
Procedure 1— Determine the nitrogen content of the protein under test as directed under Nitrogen Determination 461. Commercial instrumentation is available for the Kjeldahl nitrogen assay.
Procedure 2— Commercial instrumentation is available for nitrogen analysis. Most nitrogen analysis instruments use pyrolysis (i.e., combustion of the sample in oxygen at temperatures approaching 1000), which produces nitric oxide (NO) and similar oxides of nitrogen (NOx) from the nitrogen present in the test protein. Some instruments convert the nitric oxides to nitrogen gas, which is quantified with a thermal conductivity detector. Other instruments mix nitric oxide (NO) with ozone (O3) to produce excited nitrogen dioxide (NO2), which emits light when it decays and can be quantified with a chemiluminescence detector. A protein reference material or reference standard that is relatively pure and is similar in composition to the test proteins is used to optimize the injection and pyrolysis parameters and to evaluate consistency in the analysis.
Calculations— The protein concentration is calculated by dividing the nitrogen content of the sample by the known nitrogen content of the protein. The known nitrogen content of the protein can be determined from the chemical composition of the protein or by comparison with the nitrogen content of the USP Reference Standard or reference material.

1  Suitable standards are available from NIST (Gaithersburg, MD), Beckman Instruments (Fullerton, CA), Sigma Chemical (St. Louis, MO), Pierce (Rockford, IL), or Hewlett-Packard.
2  A suitable grade is available commercially as “Palladium Catalyst, Type I (5% Palladium on Calcium Carbonate),” from Engelhard Industries, Inc., fax number (864) 885-1375.
3  Dye purity is important in the reagent preparation. It is intended to propose a reagent footnote to indicate that Serva Blue G (Crescent Chemical Company, Hauppauge, NY) is an acceptable grade.

Auxiliary Information—
Staff Liaison : Larry N. Callahan, Ph.D., Scientist
Expert Committee : (BBPP05) Biologics and Biotechnology - Proteins and Polysaccharides
USP29–NF24 Page 2858
Pharmacopeial Forum : Volume No. 27(2) Page 2277
Phone Number : 1-301-816-8385