Fact-checked by Grok 2 weeks ago

Buffer solution

A buffer solution is a mixture of a weak acid and its conjugate base, or a weak base and its conjugate acid, that resists significant changes in when small quantities of a strong acid or strong base are added. This resistance arises because the components of the buffer neutralize added H⁺ or OH⁻ ions without substantially altering the overall hydrogen ion concentration. Buffer solutions are essential in maintaining stable levels in various chemical and biological systems. The mechanism of a buffer solution involves the between the weak acid (HA) and its conjugate base (A⁻), where added acid reacts with A⁻ to form HA, and added base reacts with HA to form A⁻ and . For example, in an acetic acid-sodium buffer, ions (CH₃COO⁻) capture protons from added HCl to produce acetic acid (CH₃COOH), while acetic acid donates protons to added NaOH to regenerate . This ensures that the remains relatively constant, with the buffer's effectiveness depending on the concentrations of its components and the pKₐ of the weak acid. The pH of a buffer solution can be precisely calculated using the Henderson-Hasselbalch equation: pH = pKₐ + log₁₀([A⁻]/[HA]), where pKₐ is the negative logarithm of the , [A⁻] is the concentration of the conjugate , and [HA] is the concentration of the weak acid. This equation allows for the design of buffers with targeted pH values by adjusting the ratio of base to acid forms. Buffer capacity, which measures the amount of acid or base a buffer can neutralize before its pH changes significantly, increases with higher concentrations of the buffering components. Buffer solutions play a critical role in biological processes by stabilizing intracellular and extracellular within narrow ranges essential for function, , and metabolic reactions. In human blood, for instance, the (HCO₃⁻/H₂CO₃) maintains around 7.4 to support and prevent or . In chemistry and industry, buffers are used in experiments, pharmaceutical formulations, and to control reactions and ensure product stability.

Fundamentals of Buffer Solutions

Definition and Composition

A buffer solution is an aqueous mixture containing a and its conjugate , or a and its conjugate acid, designed to resist substantial changes in upon the addition of small quantities of a strong acid or strong . This resistance arises from the between the acid-base pair, which allows the solution to absorb added protons or ions without drastic pH shifts. The typical composition of a buffer solution includes the weak acid or as the primary buffering agent and a that supplies the corresponding conjugate to establish the necessary . For instance, an acetate buffer comprises acetic acid (CH₃COOH) and (CH₃COONa), where the salt dissociates to provide acetate ions (CH₃COO⁻). The relative concentrations of the acid and conjugate base play a key role in setting the buffer's , with balanced ratios promoting optimal performance around the pKa value. Effective buffering requires the pKa of the weak acid (or pKb of the weak base) to be near the target , generally within one pH unit, to ensure sufficient concentrations of both species. Qualitative influences such as the solution's , which can alter ion activities, and , which affects equilibria, must also be considered for reliable performance. The term "buffer" originated in 1914, coined by G. S. Walpole to describe stabilizing mixtures in biological and chemical contexts.

Mechanism of pH Resistance

Buffer solutions resist changes in through dynamic chemical equilibria involving weak acids and their conjugate bases, or weak bases and their conjugate acids. In a weak acid buffer, such as acetic acid (HA) and (A⁻), the is established as \ce{HA ⇌ H+ + A-}. When a strong base, like (OH⁻), is added, the OH⁻ reacts with free H⁺ to form , reducing the H⁺ concentration and disturbing the . According to Le Châtelier's principle, the system responds by shifting the to the right, dissociating more HA to replenish H⁺ and thus minimizing the increase. Conversely, addition of a strong acid introduces excess H⁺, which combines with A⁻ to form undissociated HA, shifting the to the left and consuming the added H⁺ to limit the decrease. For weak base buffers, exemplified by and , the relevant is \ce{B + H2O ⇌ BH+ + OH-}. Addition of a strong acid provides H⁺ that reacts with OH⁻ to form , decreasing OH⁻ and prompting the to shift right, producing more OH⁻ from B to counteract the drop. If a strong is added, excess OH⁻ reacts with BH⁺ to form B and , shifting the left to regenerate BH⁺ and reduce the rise. This application of Le Châtelier's principle ensures that the added ions are effectively neutralized by the buffer components without significantly altering the H⁺ or OH⁻ concentrations. Buffers have inherent limitations in their pH resistance. They fail when large quantities of or are added, exceeding the available concentrations of the buffering species and depleting one component, after which the behaves like a non-buffered one with drastic pH changes. Effectiveness is also reduced if the 's pH deviates significantly from the pKa of the weak or , as the favors one form over the other, limiting the buffer's ability to absorb perturbations. Dilution with causes only minor pH shifts due to slight changes in ionic equilibria, but extreme dilution can diminish buffering capacity by reducing component concentrations.

Buffer Capacity and Effectiveness

Quantitative Definition

Buffer capacity, denoted as β, quantifies the resistance of a buffer solution to changes upon addition of or . It is defined as the number of moles of strong or strong required to alter the of one liter of the buffer solution by one unit. This measure arises from the buffer's ability to absorb added H⁺ or OH⁻ through shifts, with higher values indicating stronger buffering action. For a monoprotic buffer consisting of a weak acid HA and its conjugate base A⁻, the buffer capacity can be approximated using the Van Slyke equation derived from the Henderson-Hasselbalch relation. The Henderson-Hasselbalch equation is pH = pK_a + \log_{10} \left( \frac{[A^-]}{[HA]} \right). Differentiating with respect to added base B (where d[A^-] = -d[HA] = dB for small additions, neglecting [H⁺] and [OH⁻] contributions), yields β = \frac{dB}{dpH} \approx 2.303 \frac{[HA][A^-]}{[HA] + [A^-]}, where 2.303 is \ln(10). This approximation holds when the buffer concentration is sufficiently high relative to [H⁺] and [OH⁻]. The capacity reaches its maximum when pH = pK_a, corresponding to [HA] = [A^-] = \frac{C}{2}, where C is the total buffer concentration ([HA] + [A^-]), giving β_{max} \approx 0.576 C. Buffer capacity is typically expressed in units of moles per liter per unit (mol L⁻¹ ⁻¹). Experimentally, it is determined through : a known volume of buffer is titrated with standardized strong or while monitoring , and β is calculated as the moles of titrant added per liter divided by the observed change (often averaged over a ±0.5 range around the buffer for accuracy). A higher β signifies better stability, with the 1:1 [HA]:[A⁻] providing optimal resistance for a given concentration, though actual capacity also depends on the total buffer amount present.

Factors Affecting Capacity

The buffer capacity of a solution increases with the total concentration of the buffering components, as higher concentrations provide more weak and conjugate molecules available to neutralize added or , up to the limits imposed by and of the species. For a fixed amount of buffer solute, diluting the solution by increasing its volume reduces the concentration and thereby decreases the buffer capacity per unit volume, though the total capacity across the entire volume remains constant. Buffer capacity exhibits a strong dependence on the deviation between the solution's pH and the buffer's value, reaching a maximum at pH = and declining rapidly outside the range of pH = ± 1 unit. This relationship arises because the capacity is proportional to the product of the concentrations of the and conjugate forms, which is optimized when their ratio is 1:1. Beyond this range, the buffer becomes less effective, with capacity dropping to about half its maximum at pH = ± 1 and approaching zero farther away, as illustrated by the bell-shaped curve of buffer capacity versus , centered at the . The ratio of conjugate base to in the buffer mixture also influences , with the optimal performance occurring at a 1:1 ratio (pH = ), but buffers remain functional at extreme ratios such as 10:1 or 1:10, corresponding to pH = ± 1, where is reduced but still significant within the effective pH range. Temperature affects buffer capacity indirectly through shifts in the value, which depend on the of the reaction: endothermic dissociations increase pKa with rising , while exothermic ones decrease it. These shifts alter the pH-pKa alignment and thus the effective capacity at a given . Representative temperature coefficients (ΔpKa/ΔT) for common buffers are provided below; negative values indicate a decrease in pKa with increasing , which is typical for many biological buffers.
Buffer SystempKa (25°C)ΔpKa/ΔT (°C⁻¹)
(acetic acid)4.76-0.0002
(pK₂)7.20-0.0028
Tris (tris(hydroxymethyl)aminomethane)8.06-0.031
(pK₂)10.33-0.0096
These values highlight that buffers like acetate show minimal temperature sensitivity, while amine-based buffers like Tris exhibit pronounced shifts, impacting their suitability for temperature-variable applications. Ionic strength influences buffer capacity via the Debye-Hückel theory, which describes how increased salt concentrations alter the activity coefficients of ionic species in solution, thereby shifting the apparent and reducing the effective concentrations of active buffer forms. In high-ionic-strength environments, such as those with added salts exceeding 0.1 M, this effect diminishes capacity by up to 10-20% for monoprotic buffers, particularly those involving charged species, necessitating adjustments in formulation to maintain performance.

Types of Buffer Solutions

Simple Buffering Agents

Simple buffering agents are solutions composed of a single and its conjugate , or a and its conjugate acid, which together resist changes upon addition of small amounts of acid or base. These systems rely on the equilibrium between the acid and its salt to maintain stability, typically effective within approximately 1 pH unit of the acid's value. Preparation of simple buffers commonly involves dissolving equimolar amounts of the weak and its conjugate salt in water to achieve a pH near the pKa. Alternatively, partial neutralization of the weak with a strong (or vice versa) can produce the desired ratio of to conjugate, or existing solutions can be adjusted by adding small volumes of strong or to fine-tune the pH. Stock solutions of these components are often prepared separately and mixed to create working buffers, ensuring consistency and ease of use in settings. Common examples include the acetate buffer, formed from acetic acid (pKa 4.76 at 25°C) and , which operates effectively in the range of 3.6 to 5.6. Citrate buffers, using (pKa values 3.13, 4.76, and 6.40 at 25°C) in a simple pair such as the first dissociation with its monosodium salt, provide buffering around 3 to 5 or 4 to 6 depending on the selected pair. Borate buffers, derived from (pKa 9.24 at 25°C) and , are suitable for alkaline conditions in the range of 8 to 10. These agents offer advantages such as straightforward preparation and low cost, making them accessible for routine applications. However, their buffering range is inherently narrow, limited to about ±1 unit from the , and certain options like buffers carry potential toxicity risks, including reproductive harm from exposure.

Universal Buffer Mixtures

Universal buffer mixtures are formulations combining multiple buffering agents to provide effective pH control across a broad range, typically from pH 2 to 12, allowing researchers to maintain consistent ionic environments without preparing entirely new solutions for different pH values. These systems rely on the overlapping values of the constituent weak acids and their conjugates, enabling sequential dominance of buffering action as pH changes. Prominent examples include the Britton-Robinson buffer, developed in 1931, which consists of 0.04 M each of acetic acid, orthophosphoric acid, and boric acid, with the pH adjusted using 0.2 M sodium hydroxide to cover the range from pH 2 to 12. Another widely used formulation is the McIlvaine buffer, introduced in 1921, comprising 0.1 M citric acid and 0.2 M disodium hydrogen phosphate mixed in varying proportions to achieve pH values from 2.2 to 8.0. Phosphate-based variants attributed to Sørensen, such as the standard 0.067 M sodium phosphate buffer (combining Na₂HPO₄ and KH₂PO₄), provide coverage in the narrower range of pH 5.8 to 8.0 but can be modified with additional components for extended utility in universal applications. Preparation of these mixtures typically involves dissolving the acid components in deionized water to form a stock solution, followed by stepwise addition of a strong base like NaOH while monitoring with a calibrated , ensuring gradual to avoid overshooting. For instance, in the Britton-Robinson system, the acids are combined first, then NaOH is added incrementally until the target is reached, with final volume adjustment using water. Stability concerns arise during this process, including potential precipitation of borates or phosphates at extreme values, which can occur if is not controlled or if incompatible ions are present, necessitating careful storage at and use within weeks to minimize degradation. The primary advantage of universal buffer mixtures is their versatility, enabling experiments across wide pH ranges with a single base formulation, which simplifies workflows in electrochemical, spectroscopic, and studies. However, this comes at the cost of increased complexity in preparation compared to single-component buffers, potential chemical interactions between agents that may alter or introduce artifacts, and reduced buffering capacity per pH unit due to the distributed concentrations of individual components.

Biological Buffer Systems

Biological buffer systems encompass a range of buffering agents specifically developed or selected for compatibility with living organisms and biochemical processes, ensuring minimal interference with cellular functions while maintaining stable in physiological ranges. These buffers are essential in involving enzymes, proteins, and cultures, where pH fluctuations can denature biomolecules or disrupt metabolic pathways. Unlike general-purpose buffers, biological ones prioritize properties that support , such as in aqueous media and low to cells. The development of modern biological buffers traces back to the 1960s, when biochemist Norman Good and colleagues introduced a series of zwitterionic compounds known as to address limitations of traditional agents like or in biochemical experiments. These buffers were engineered for use in studies of proteins and cellular organelles, offering enhanced stability and reduced interaction with biological components compared to earlier options. Their zwitterionic structure—featuring both positive and negative charges—contributes to high and ionic strength mimicry of physiological conditions, making them staples in research. Subsequent refinements have expanded the lineup, but Good's original set remains foundational for life sciences applications. Selection criteria for biological buffers emphasize attributes that prevent adverse effects on experimental systems, including a pKa value between 6.0 and 8.0 to align with most cellular pH optima, high solubility, and low permeability through membranes to avoid unintended shifts. Additional requirements include minimal of metal ions, which could disrupt cofactors; low absorbance in the (UV) and visible spectra (typically below 230–700 nm) to enable spectroscopic assays without interference; and chemical stability under physiological temperatures and ionic conditions. Buffers must also exhibit non-toxicity, prompting avoidance of agents like in due to potential metabolic interference or . These criteria ensure buffers support rather than hinder biological integrity. Common biological buffers include several Good's variants alongside established inorganic options, each tailored to specific pH needs and experimental contexts. The following table summarizes key examples:
Buffer NamepKa (at 20–25°C)Effective pH RangeNotable Properties and Uses
Tris (tris(hydroxymethyl)aminomethane)8.06–8.17.0–9.0Temperature-sensitive (pH decreases ~0.03 units/°C); widely used in protein electrophoresis and enzyme assays due to its solubility and compatibility with biomolecules, though its basic range limits lower-pH applications.
HEPES (4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid)7.56.8–8.2A Good's buffer with low UV absorbance and minimal metal chelation; ideal for cell culture and mammalian systems as it mimics physiological ionic environments without toxicity.
Phosphate (e.g., Na2HPO4/NaH2PO4)7.2 (pKa2)5.8–8.0Inorganic buffer with high biocompatibility and low cost; effective in isotonic solutions for maintaining extracellular pH, though it can precipitate with divalent cations at higher concentrations.
MOPS (3-(N-morpholino)propanesulfonic acid)7.2–7.286.5–7.9Good's buffer featuring a morpholine ring for stability; suitable for near-neutral pH in cell lysis, protein purification, and media formulation due to its low salt effects and non-interference with UV detection.
In practice, (), a mixture of salts with and at 7.4, serves as a versatile for washing cells, diluting samples, and transporting tissues in biological workflows, preserving cell viability without osmotic stress. buffers, often as systems equilibrated with CO2, are employed in incubators to simulate physiological and stabilize in media for CO2-dependent organisms. These examples highlight how biological buffers integrate seamlessly into research protocols to replicate conditions .

Applications of Buffer Solutions

Laboratory and Analytical Chemistry

In laboratory and analytical chemistry, buffer solutions play a crucial role in calibrating pH measurement instruments to ensure accurate and reproducible results. Standard buffer solutions, such as those developed by the National Institute of Standards and Technology (NIST), are used to calibrate pH meters by providing reference points at specific pH values. For instance, potassium hydrogen phthalate buffer (SRM 185i) yields pH 4.005 at 25°C, phosphate buffer (SRM 186g) provides pH 6.864 (equimolal formulation) or 7.416 (physiological formulation) at 25°C, and sodium carbonate/sodium bicarbonate buffer (SRM 191d) achieves pH 10.014 at 25°C, all traceable to primary standards for high precision within ±0.01 pH units. These buffers are prepared from high-purity reagents and certified against electrometric methods to minimize variability during calibration of glass electrodes in electrometric pH assemblies. Buffer solutions are essential in enzymatic assays to maintain a stable environment that matches the optimal conditions for activity, thereby ensuring reliable kinetic measurements and reaction outcomes. For example, buffers at pH 4.5 are commonly employed in assays involving (GUS) from sources like or bovine liver, where they facilitate reactions by resisting pH shifts caused by proton release or addition. This stability is critical for , as even minor pH fluctuations can alter conformation and activity rates, leading to inaccurate results in biochemical studies. In chromatographic techniques, buffers control the and of mobile phases to influence retention and separation efficiency, particularly in (HPLC) and ion-exchange chromatography. Phosphate buffers, often at concentrations of 10–50 mM and 6–8, are widely used in reversed-phase HPLC and ion-exchange methods for protein separation, as they provide effective buffering capacity while being compatible with UV detection and helping to modulate protein charge for selective and . Quality control is paramount for buffer solutions in analytical settings to prevent errors from degradation or impurities. Certified buffers, such as those compliant with () standards under <791> , must be fresh and stored in tightly sealed, alkali-free glass containers to avoid from CO₂ absorption or microbial . Buffers should be used within their , typically 1-2 years unopened when stored properly (10–25°C, protected from light). After opening, replace every 1-3 months or sooner if signs of degradation appear, per <791> guidelines. verification is often conducted by ISO 17025-accredited labs against NIST references.

Industrial and Pharmaceutical Uses

In pharmaceutical formulations, buffers play a critical role in stabilizing for injectables and oral to ensure and prevent degradation. Citrate buffers, for instance, are commonly employed in formulations to maintain an optimal acidic range, facilitating the disaggregation of viral particles and enhancing stability during storage and administration. As of 2025, citrate buffers are increasingly used in mRNA-lipid (LNP) formulations to optimize stability and encapsulation efficiency at 3.0-6.0. is frequently used as an in various to control and support overall integrity. Similarly, tartrate salts, such as in phenindamine , serve as buffering agents in formulations, aiding stabilization in oral preparations to improve and shelf-life. buffers are also utilized in pharmaceutical preparations for their compatibility in maintaining during freeze-drying processes. In biotechnological production, buffers are essential for controlling during processes, particularly in the large-scale synthesis of . and ammonium-based buffers, such as mixtures of and , are added to maintain a stable around 6.8, preventing fluctuations that could inhibit microbial growth and reduce yields in processes like L-lysine or L-glutamic acid production. These buffers counteract the natural rise in pH due to liberation from , ensuring consistent conditions in industrial bioreactors. Buffers are integral to for precise adjustment in processes like textile and . In operations, acidic buffers such as or citrate systems are used to stabilize between 4 and 6, promoting uniform uptake on fibers and minimizing color variations. For , serves as a key buffering agent in nickel electrolytes to keep steady at 4–4.5 near the , preventing hydrogen evolution and ensuring even metal deposition. Regulatory frameworks from the FDA and outline guidelines for buffer excipients to guarantee safety and consistency in pharmaceutical products. The FDA's guidance on nonclinical studies for pharmaceutical excipients requires toxicity assessments for buffers like citrate and tartrate to confirm their suitability in drug formulations, emphasizing compatibility with active ingredients. Similarly, aligns with ICH Q1A(R2) stability testing protocols, which mandate evaluating buffer performance under accelerated conditions (e.g., 40°C/75% RH) to verify maintenance and product stability over . These standards ensure buffers meet pharmacopeial requirements for purity and functionality in commercial manufacturing.

Physiological and Environmental Roles

In the , the plays a central role in maintaining blood at approximately 7.4 through the between (CO₂), , (H₂CO₃), and ions (HCO₃⁻), which allows rapid adjustment via respiratory and renal mechanisms. This system resists pH fluctuations from metabolic acids, with the ratio of HCO₃⁻ to H₂CO₃ typically around 20:1 under normal conditions. buffers contribute significantly in , where the pKₐ of 6.8 enables H₂PO₄⁻ to capture excess protons during acid , preventing drastic pH drops in renal tubular fluid. Proteins, including plasma albumins and intracellular enzymes, further enhance buffering through ionizable groups on amino acid side chains, such as residues, which donate or accept protons near physiological pH. Hemoglobin serves as a key buffer in blood, particularly in red blood cells, where its histidine-rich structure (with pKₐ values around 6.6–7.85) binds protons released during CO₂ transport, stabilizing pH during oxygen delivery and preventing acidosis in tissues. In cellular environments across organisms, amino acid side chains—especially from , aspartate, and glutamate—act as intracellular buffers by ionizing in response to pH changes, maintaining optimal conditions for enzymatic activity and metabolic processes. These protein-based systems are ubiquitous in vertebrates and , underscoring their evolutionary importance for acid-base . In natural environments, the ocean's buffer system regulates at about 8.1 via the equilibrium of CO₂, , and ions, enabling the absorption of atmospheric CO₂ while supporting calcifying like corals and . However, ongoing —driven by rising CO₂ levels—has lowered average surface by 0.1 units since pre-industrial times, reducing buffer capacity and threatening marine ecosystems through impaired shell formation and . In soils, clay minerals and humic acids provide buffering against shifts from rainfall or fertilization; clays exchange cations to neutralize acidity, while , with their carboxylic and groups, bind protons and stabilize soil for plant growth. Disruptions to these natural buffers can lead to significant imbalances; in humans, like (excess acid from or ) or ( excess from ) overwhelm buffer systems, causing deviations that impair organ function and require interventions like . Environmentally, —laden with sulfuric and nitric acids—depletes lake buffering capacity in low-alkalinity watersheds, lowering below 5 in sensitive areas and harming populations through aluminum mobilization and reproductive failure.

pH Calculations for Buffers

Monoprotic Buffer Systems

Monoprotic buffer systems consist of a (HA) and its conjugate base (A⁻), or a and its conjugate acid, where only one proton occurs. The of such buffers is calculated using the Henderson-Hasselbalch equation, which relates the to the (pKₐ) and the ratio of conjugate base to acid concentrations. This equation is derived from the equilibrium expression for the weak acid . The acid dissociation constant K_a for HA ⇌ H⁺ + A⁻ is given by K_a = \frac{[H^+][A^-]}{[HA]} Rearranging yields [H⁺] = K_a \frac{[HA]}{[A^-]}. Taking the negative logarithm of both sides of the K_a expression gives -\log K_a = -\log [H^+] + \log \frac{[A^-]}{[HA]} which simplifies to pK_a = pH + \log \frac{[HA]}{[A^-]} or, equivalently, pH = pK_a + \log \frac{[A^-]}{[HA]}. This form, known as the Henderson-Hasselbalch equation, was first proposed by Lawrence J. Henderson in 1908 and reformulated in logarithmic terms by Karl A. Hasselbalch in 1916. The equation enables prediction of buffer pH from known concentrations or ratios of HA and A⁻. For instance, when [A⁻]/[HA] = 1, pH = pKₐ, indicating maximum buffering capacity at this ratio. To achieve a target pH, the required ratio is [A⁻]/[HA] = 10^(pH - pKₐ); for example, a ratio of 10 yields pH = pKₐ + 1, and 0.1 yields pH = pKₐ - 1. These applications assume the buffer operates near the pKₐ for effective resistance to pH changes. The derivation relies on key assumptions: the dissociation of HA and hydrolysis of A⁻ are negligible compared to their initial concentrations, valid in dilute solutions where buffer concentrations exceed [H⁺] and [OH⁻] by at least 100-fold; the pH lies within pKₐ ± 1; and concentrations approximate activities. Errors arise from non-ideal behavior, such as ionic strength effects on activity coefficients, which deviate from ideality in concentrated or high-ionic-strength solutions, potentially leading to pH inaccuracies of up to 0.3 units. The equation does not apply to strong acids or bases, where dissociation is complete. A representative example is an acetate buffer with 0.10 M acetic acid (CH₃COOH, pKₐ = 4.76) and 0.10 M (CH₃COONa), where [A⁻]/[HA] = 1. Substituting into the equation gives pH = 4.76 + (1) = 4.76, demonstrating the buffer's equals the pKₐ at equal concentrations.

Polyprotic Buffer Systems

Polyprotic buffer systems involve acids or bases capable of donating or accepting multiple protons, leading to more complex calculations compared to monoprotic systems due to successive steps. For a diprotic acid H₂A, such as (H₃PO₄, where the relevant steps are H₃PO₄/H₂PO₄⁻ and H₂PO₄⁻/HPO₄²⁻), the is approximated using the Henderson-Hasselbalch equation for the dominant conjugate pair when the target lies between pK_{a1} and pK_{a2}. In this range, the first dominates, and the equation simplifies to: \text{pH} \approx \text{p}K_{a1} + \log \left( \frac{[\ce{HA-}]}{[\ce{H2A}]} \right) where [HA⁻] and [H₂A] are the concentrations of the intermediate and fully protonated forms, respectively. This approximation holds because the second dissociation contributes negligibly to [H⁺] when K_{a1} ≫ K_{a2}, allowing treatment of the system as effectively monoprotic for that step. For more precise calculations, especially when adding to H₂A, the full system must be considered, incorporating both constants. The charge balance and equations lead to a in [H⁺], but successive approximations are often used: first solve for the first ignoring the second, then refine by including contributions from subsequent steps. Iterative numerical methods, such as successive substitution, converge to the exact by updating concentrations until stability. In cases of overlapping pK_a values (where ΔpK_a < 3), these approximations break down, requiring computational tools like software (e.g., Visual MINTEQ or PHREEQC) to solve the coupled equilibria accurately. A key example is the carbonic acid-bicarbonate buffer in , where H₂CO₃ (pK_{a1} = 6.35) dissociates to HCO₃⁻ and the second step to CO₃²⁻ (pK_{a2} = 10.33) is minimal at physiological pH ~7.4. Here, the dominant species are H₂CO₃ and HCO₃⁻, so pH ≈ 6.35 + log([HCO₃⁻]/[H₂CO₃]), with typical ratios maintaining pH despite CO₂ variations. Another common system is (pK_{a1} = 3.13, pK_{a2} = 4.76, pK_{a3} = 6.40), used in buffers for pH 3–6; buffers around pK_{a2} or pK_{a3} employ the intermediate form H Cit²⁻ as the dominant species, acting amphoterically (as both and ) to resist pH shifts. In phosphate buffers, the amphoteric HPO₄²⁻ species (from pK_{a2} = 7.20 of ) similarly facilitates buffering near neutral pH by participating in both proton donation and acceptance.

References

  1. [1]
    Buffer Solutions
    A buffer solution is one in which the pH of the solution is "resistant" to small additions of either a strong acid or strong base. Buffers usually consist of a ...Missing: definition | Show results with:definition
  2. [2]
    15.6 Buffers – Chemistry Fundamentals - UCF Pressbooks
    A solution containing appreciable amounts of a weak conjugate acid-base pair is called a buffer solution, or a buffer. Buffer solutions resist a change in pH ...
  3. [3]
    How Buffers Work - ChemCollective
    If a strong base is added to a buffer, the weak acid will give up its H + in order to transform the base (OH - ) into water (H 2 O) and the conjugate base.Missing: mechanism | Show results with:mechanism
  4. [4]
    [PDF] Buffers | Calbiochem
    It is a measure of the protection a buffer offers against changes in pH. Buffer capacity generally depends on the concentration of buffer solution. Buffers ...
  5. [5]
    Buffers
    In effect, a buffer solution behaves somewhat like a sponge that can absorb H + and OH − ions, thereby preventing large changes in pH when appreciable amounts ...<|control11|><|separator|>
  6. [6]
    Henderson-Hasselbach
    This is the well-known Henderson-Hasselbalch equation that is often used to perform the calculations required in preparation of buffers for use in the ...
  7. [7]
    [PDF] Experiment # 9: The Henderson-Hasselbalch Equation - ULM
    A buffer solution actually is a mixture of a weak acid and its conjugate base or a mixture of a weak base and its conjugate acid. The conjugate forms are ...
  8. [8]
    Lecture 22: Acid-Base Equilibrium: Salt Solutions and Buffers
    Our blood has a natural buffering system to ensure that the pH of our blood stays within a narrow window and that we stay health.
  9. [9]
    [PDF] EXPERIMENT 6 - Buffer Effects - Chemistry at URI
    Buffers are solutions that contain an acid and its conjugate base that are designed to resist pH changes. This is important in biological systems to maintain ...
  10. [10]
    The Old West analogy for acid-base buffering - PMC - NIH
    A buffer is a mixture of a weak acid and its conjugate base (eg, carbonic acid and bicarbonate), or a mixture of a weak base and its conjugate acid.
  11. [11]
    8.8: Buffers – CHM130 Fundamental Chemistry
    Key Takeaway. A buffer is a solution that resists sudden changes in pH. A buffer consists of a conjugate weak acid-weak base pair.
  12. [12]
    [PDF] Making Buffers
    Choose an acid whose pKa is closest to the desired buffer pH. Buffers are usually made from acid/conjugate base systems that have pKa = desired pH ± 1.
  13. [13]
    [PDF] Buffer Solutions - CHEMISTRY 1B
    Buffer Solutions. Introduction. Any solution that contains both a weak acid HA and its conjugate base A. – in significant amounts is a buffer solution. A ...Missing: composition | Show results with:composition
  14. [14]
    17.2: Buffers- Solutions That Resist pH Change - Chemistry LibreTexts
    Aug 11, 2022 · Le Châtelier's Principle states that if an equilibrium becomes unbalanced, the reaction will shift to restore the balance. If a common ion is ...
  15. [15]
    3.2: Buffer Solutions - Chemistry LibreTexts
    May 8, 2021 · Again, this is consistent with Le Chatelier's principle: adding H ⁡ A + ions drives the dissociation equilibrium to the left. Exercise ...
  16. [16]
    7.2: Practical Aspects of Buffers - Chemistry LibreTexts
    Oct 23, 2023 · Buffer solutions do not have an unlimited capacity to keep the pH relatively constant (Figure 7 . 2 . 1 ). If we add so much base to a buffer ...
  17. [17]
    Buffers and buffering power - Deranged Physiology
    Dec 18, 2023 · Conversely, when the buffer is at a pH which is far from its pKa, it becomes much less capable as a buffer. If only 1% of the buffer is ionised, ...
  18. [18]
    17.2: Buffer Solutions - Chemistry LibreTexts
    Jul 12, 2023 · Buffers are solutions that maintain a relatively constant pH when an acid or a base is added. They therefore protect, or “buffer,” other molecules in solution.
  19. [19]
    (PDF) Understanding, Deriving, and Computing Buffer Capacity
    Aug 6, 2025 · The β values were calculated by the Van Slyke equation [28] using the pK a values close to pH 6.5 (BCB: pK a = 6.05 (for I = 0.15 M, at 37°C) ...
  20. [20]
    Buffer solutions in drug formulation and processing: How pKa values ...
    Apr 5, 2019 · The pK a of buffers depends on temperature, pressure and ionic strength. In addition, the pK a can also be affected by the polarity of the solvent.
  21. [21]
    Dissociation constant pKa (25 °C), dpKa/dT, and molecular mass of ...
    Jul 2, 2019 · Dissociation constant pKa (25 °C), dpKa/dT, and molecular mass of buffers ; Phosphate (phosphoric acid) (pK1), 2.15, 0.0044 ; Glycine (pK1), 2.35 ...
  22. [22]
    Development of Methods for the Determination of pKa Values - PMC
    Aug 8, 2013 · Activity coefficients (γ) of different ionic species in solution strongly depend on ionic strength (Debye-Hückel theory). Because pKa depends on ...
  23. [23]
    7.1: Acid-Base Buffers - Chemistry LibreTexts
    Oct 22, 2023 · A solution containing a mixture of an acid and its conjugate base, or of a base and its conjugate acid, is called a buffer solution.
  24. [24]
    Methods for preparing buffers (video) | Khan Academy
    Nov 4, 2021 · In this video, we'll explore two common methods for preparing buffer solutions. In the first approach, a certain amount of a weak acid (or weak base) is ...
  25. [25]
    Buffer Preparation – Solutions, Calculation & Solving Common ...
    A buffer keeps the pH of a solution constant by absorbing protons that are released during reactions or by releasing protons when they are consumed by reactions ...Missing: mechanism | Show results with:mechanism<|control11|><|separator|>
  26. [26]
    Appendix C: Dissociation Constants and pKa Values for Acids at 25°C
    Name, Formula, K a1, pK a1, K a2, pK a2, K a3, pK a3, K a4, pK a4. Acetic acid, CH 3CO 2H, 1.75 × 10 −5, 4.756. Arsenic acid, H 3AsO 4, 5.5 × 10 −3, 2.26 ...
  27. [27]
    Acetate Buffer (pH 3.6 to 5.6) Preparation and Recipe | AAT Bioquest
    Acetate buffers are inexpensive and simple to prepare, and can be stored at room temperature. To prepare. L of Acetate Buffer (pH 3.6 to 5.6): Change the value ...
  28. [28]
    Boric Acid | H3BO3 | CID 7628 - PubChem
    The pKa of boric acid is 9.24(3), indicating that this compound will exist primarily in the undissociated form in the environment, but partially in the ...
  29. [29]
    Toxicity of boric acid, borax and other boron containing compounds
    Jan 22, 2021 · Skin exposure to boric acid has proven fatal in some cases, and the range of toxicity effects include abdominal as well as local effects on the skin.Missing: buffer | Show results with:buffer
  30. [30]
    Universal buffer solutions and the dissociation constant of veronal
    —Universal buffer solutions and the dissociation constant of veronal. H. T. S. Britton and R. A. Robinson, J. Chem. Soc., 1931, 1456 DOI: 10.1039 ...Bevat niet: formulations McIlvaine Sørensen scientific
  31. [31]
    Universal buffers for use in biochemistry and biophysical experiments
    One of the most common buffering agents is phosphate-buffered saline (PBS) which was formulated to match the ionic strength and pH of mammalian cells.
  32. [32]
    [PDF] pH values of the Clark and Lubs buffer solutions at 25° C
    The pH valucs of the well-known Clark and Lubs burrcr solutions havc been dctermined at 25° C on the convcntional activity pH scale dcfined by thc NBS ...
  33. [33]
    Preparing Buffer Solutions - Shimadzu
    Good separation reproducibility (stability) may not be achieved if buffer solutions are not used. A buffer solution is prepared as a combination of weak acids ...Missing: universal | Show results with:universal
  34. [34]
    A New extractive spectrophotometric method for determination of ...
    The pH value was adjusted to 3.0. The Britton- Robinson buffers in the pH range of 2–7 were prepared from an equal mixture of 0.1 M acetic acid, 0.1 M boric ...
  35. [35]
    HPLC Troubleshooting: Precipitation of Buffers in Gradient Elution ...
    Mar 14, 2022 · Buffer precipitation occurs with phosphate/biological buffers in high organic content, especially when mixing with 100% organic solvents, and ...Missing: universal stability
  36. [36]
    Buffer Solution - an overview | ScienceDirect Topics
    In measurements by ion-selective electrodes, the term ionic strength buffer is used to denote a solution that is added to keep the ionic strength, and in ...Missing: distinction | Show results with:distinction
  37. [37]
    [PDF] Good's buffers (biological buffers) - Interchim
    These so-called “Good's Buffers” are widely used in cell culture and other biological applications. Since then, additional zwitterionic buffers (AMPSO, CAPSO, ...
  38. [38]
  39. [39]
  40. [40]
    [PDF] Biological Buffers and Ultra Pure Reagents
    These so-called “Good's Buffers” are now widely used in cell culture, electrophoresis, biological systems and biochemical reactions. Over the years, several new ...
  41. [41]
    What Makes a “Good” Laboratory Buffer? - Bitesize Bio
    May 29, 2025 · Most biological reactions take place at a pH between 6 and 8, so ideal buffers have pKa values in this range to provide maximum buffering capacity there.Missing: applications | Show results with:applications<|control11|><|separator|>
  42. [42]
    Overview: Good's Buffers - Boston BioProducts
    Low metal chelating capability; Have minimal salt effects. Have minimal ... Have low optical absorbance either in the visible or in the UV region. Upon ...
  43. [43]
    Tris Buffer | Bio-Rad
    Making a Tris Buffer​​ Tris buffer is a good choice for most biological systems because it has a pKa of approximately 8.1 at 25°C, making it an effective buffer ...
  44. [44]
    Tris Buffer Solutions | Thermo Fisher Scientific
    Used as a biological buffer substance in the pH range 7—9; pKa = 8.3 at 20°C; pKa = 7.82 at 37°C. Used as a biological buffer or as a ...
  45. [45]
    HEPES, TRIS buffer and pH control - Blog - Hopax Fine Chemicals
    Jan 17, 2024 · Tris with a pKa around 8.1, performs exceptionally well in a higher pH range, particularly in studies involving proteins and enzymes. Effective ...
  46. [46]
  47. [47]
    HEPES Buffers - Boston BioProducts
    With a pKa of 7.5 at 25°C, HEPES effectively buffers in the pH range of 6.8 to 8.2, making it an excellent choice for cell culture, enzyme ...
  48. [48]
    Phosphate Buffer (pH 5.8 to 7.4) Preparation and Recipe
    A simple phosphate buffer is used ubiquitously in biological experiments, as it can be adapted to a variety of pH levels, including isotonic.
  49. [49]
    Buffers for Biochemical Reactions - Promega Corporation
    All you need to know about preparing buffers, including what makes a good buffer, how buffers work, and how to prepare a buffer solution.Missing: universal | Show results with:universal
  50. [50]
    MOPS Buffer (1 M, pH 7.4) - Boston BioProducts
    With a pKa of 7.20, MOPS is an excellent buffer for many biological systems at near-neutral pH, and maintains a pH range from 6.5 to 7.9.
  51. [51]
    PBS (Phosphate Buffered Saline) (1X, pH 7.4) Preparation and Recipe
    PBS is an isotonic buffer frequently used in biological applications, such as washing cells, transportation of tissues, and dilutions.
  52. [52]
  53. [53]
  54. [54]
    pH Metrology | NIST - National Institute of Standards and Technology
    NIST pH SRMs serve as benchmarks for pH measurements in all fields of industry, government, and academia.
  55. [55]
    [PDF] Standard Reference Material® 186g - Certificate of Analysis
    Certified Values and Uncertainties: The certified pH(S) values provided in Tables 1 and 2 correspond to log (1/aH), where aH is the conventional activity of the ...
  56. [56]
    [PDF] Provisional pH values for certain standard buffer solutions
    For use in the calibration of electrometric pH assemblies, 17 standard buffer solutions have been investigated, and pH values at 20°, 25°. and 30° C have.
  57. [57]
  58. [58]
  59. [59]
    Ready-To-Use pH Buffer Solutions according to EP and USP | Merck
    Discover Certipur® certified ready-to-use pH buffer solutions acc. to EP and USP. No preparation. Just open the pack for fast, exact pH calibration.
  60. [60]
    Citrate-mediated disaggregation of rotavirus particles in RotaTeq ...
    Citrate in the formulation buffer facilitates the disaggregation of viral particles, likely through a carboxylic-acid mediated interaction.
  61. [61]
    Vaccine ingredients
    May 26, 2022 · Polysorbate 80 and sodium citrate are commonly used in food and drink. Sorbitan trioleate is a compound made from oleic acid (a natural fatty ...
  62. [62]
    Phenindamine Tartrate | C23H25NO6 | CID 11290 - PubChem
    Phenindamine Tartrate | C23H25NO6 | CID 11290 - structure, chemical names ... Drugs that selectively bind to but do not activate histamine H1 receptors, thereby ...
  63. [63]
    Thermophysical Properties of Pharmaceutically Compatible Buffers ...
    Aug 5, 2025 · Sodium tartrate, sodium malate, potassium citrate, and sodium citrate buffers were prepared with a pH range within their individual buffering ...Missing: antihistamines | Show results with:antihistamines
  64. [64]
    US5763230A - Amino acid fermentation processes - Google Patents
    When pH control at the set-point of 6.8 is done using a mixture of ammonia and ammonium sulphate in a 1:0.8 ratio (based on equimolar nitrogen contents of the ...
  65. [65]
    [PDF] pH Control In Fermentation Plants | Emerson
    The liberation and accumulation of ammonia from the metabolism of amino compounds will cause the pH to slowly rise. The pH is allowed to rise to about 7 and is ...
  66. [66]
    Why Acid Buffers are Important in Textile Dyeing and Finishing
    An acid buffer is a pH-stabilizing agent used in textile dyeing to maintain optimal acidity (pH 4–6). It ensures even dye absorption, prevents color variation, ...
  67. [67]
    How to Maximize Nickel Plating Performance | Technic Inc.
    The standard pH for a nickel electrolyte is around 4 – 4.5. To understand why we need a buffer, we need to know what happens on the surface of the cathode.
  68. [68]
    [PDF] Guidance for Industry - FDA
    This guidance describes the types of toxicity data that the Agency uses in determining whether a potential new excipient is safe for use in human ...
  69. [69]
    Buffers in Biologics Manufacturing - BioProcess International
    Feb 13, 2017 · For formulation and storage, buffers ensure optimal biological activity and stability of a drug product at the targeted storage temperature ...
  70. [70]
    Physiology, Acid Base Balance - StatPearls - NCBI Bookshelf
    Carbonic acid then dissociates into bicarbonate and a hydrogen ion. This reaction is 1 of the many buffer systems in the human body; it resists dramatic changes ...
  71. [71]
    Acid–base balance: a review of normal physiology - PMC - NIH
    From the buffer titration curve, we see that this is well below the physiological pH; at pH 7.4 the ratio of HCO3– to H2CO3 is 5000:1.
  72. [72]
    Fluid, Electrolyte, and Acid-Base Balance - OERTX
    As with the phosphate buffer, a weak acid or weak base captures the free ions, and a significant change in pH is prevented. Bicarbonate ions and carbonic acid ...<|separator|>
  73. [73]
    Blood buffers: The viewpoint of a biochemist - PMC - NIH
    May 5, 2025 · Buffer capacity (in mEq/L of whole blood) was calculated using the equation: β = 2.3 × Ka × C × [H3O+]/(Ka + [H3O+])2 (Van Slyke, 1922). The ...
  74. [74]
    [PDF] 1: Amino Acids and the Role of pH
    The side chains of the basic amino acids accept protons (see Fig. ... All free amino acids, plus charged amino acids in peptide chains, can serve as buffers.
  75. [75]
    Ocean acidification | National Oceanic and Atmospheric Administration
    Sep 25, 2025 · The ocean's average pH is now around 8.1 , which is basic (or alkaline), but as the ocean continues to absorb more CO2, the pH decreases and ...
  76. [76]
    Ocean Acidification - Woods Hole Oceanographic Institution
    Ocean acidification is a reduction in the pH of the ocean ... Since the dawn of industrialization, average surface ocean pH has decreased to about 8.1.
  77. [77]
    Soil Acidity and Liming - CTAHR
    Thus, the hydrogen ions retained by clay particles replenish, or buffer, the hydrogen ions in soil water. Table 1. pH of some common items.
  78. [78]
    Disorders of Acid-Base Balance: New Perspectives - PMC
    Dec 10, 2016 · Disorders of acid-base involve the complex interplay of many organ systems including brain, lungs, kidney, and liver.
  79. [79]
    Effects of Acid Rain | US EPA
    Mar 19, 2025 · Buffering Capacity​​ Many forests, streams, and lakes that experience acid rain don't suffer effects because the soil in those areas can buffer ...
  80. [80]
    10.3 Acid Rain – Introduction to Environmental Sciences and ...
    Lakes and streams become acidic (i.e., the pH value goes down) when the water and surrounding soil cannot buffer the acid rain enough to neutralize it.
  81. [81]
    The Henderson-Hasselbalch Equation: Its History and Limitations
    This article examines the evolution of the Henderson-Hasselbalch equation and presents a critical evaluation of its usefulness.
  82. [82]
    Lecture 3: Acid-Base Equilibria & Buffers. - andrew.cmu.ed
    Stronger acids have larger Ka values since they are more fully dissociated. The Henderson-Hasselbalch ... Acetic Acid. 4.76. Weak. Ammonium (NH4+). 9.25. Very ...
  83. [83]
    [PDF] Buffers & Polyprotic Acids f [A-] = R (1+ R) R - andrew.cmu.ed
    Sep 7, 2005 · 4.1 Solving pH Problems: The H&H equation has predictive value. Once you know the pKa of an acid and the pH, you can predict [A-]/[HA].Missing: approximations | Show results with:approximations
  84. [84]
    [PDF] Chapter 11 – Buffers, pH curves, Titrations, and Indicators
    If you make a buffer with a polyprotic acid/base you need to know WHICH Ka to use. Answer, look at which conjugate pairs are the dominant species. Once you ...
  85. [85]
    Calculating approximate pH of polyprotic acids
    Feb 3, 2014 · You get a "higher order" Henderson-Hasselbach equation: pH=(pKa1+pKa2+pKa3−log[PO3−4][H3PO4])/3. Assuming the ion concentrations are ...
  86. [86]
    Buffer Calculator
    The pKa's for phosphoric acid are 2.15, 7.20, and 12.38 at 25°C. Buffers made with the above salts work best in the pH range 6-10. What buffer strength to use?