Fact-checked by Grok 2 weeks ago

Molality

Molality is a measure of concentration in solutions, defined as the amount of substance (in moles) of solute per kilogram of solvent. It is denoted by the symbol m and expressed in units of mol/kg, providing a temperature-independent alternative to other concentration units. Unlike molarity, which is based on the volume of the entire solution and varies with temperature due to changes in density, molality relies solely on the mass of the solvent, making it constant regardless of temperature fluctuations. This property renders molality particularly useful in physical chemistry calculations where thermal effects might otherwise complicate measurements. Molality plays a central role in the study of , such as , , vapor pressure lowering, and , which depend on the number of solute particles rather than their nature. For instance, the of a solution is given by ΔTb = Kb × m, where Kb is the ebullioscopic constant of the solvent, and similarly for with ΔTf = Kf × m. These applications are essential in fields like , , and industrial processes involving solutions.

Fundamentals

Definition

Molality, denoted by the symbol m, is defined as the (typically in moles) of a solute divided by the of the in kilograms. This concentration measure focuses specifically on the solvent's as the denominator, providing a ratio that quantifies the solute's dispersion relative to the solvent's weight. The general formula for molality is m = \frac{n_{\text{solute}}}{m_{\text{solvent}}} where n_{\text{solute}} represents the number of moles of the solute and m_{\text{solvent}} is the of the in kilograms. By using rather than for the solvent, molality remains invariant with and changes, unlike volume-dependent measures where or contraction alters the solution's . For example, consider a binary aqueous solution where 58.44 g (1.0 mol) of (NaCl) is dissolved in 1.0 of ; the molality is calculated as m = 1.0 / 1.0 = 1.0 mol/, often denoted simply as 1.0 m. In SI units, molality is expressed as moles per ().

Historical Origin

The concept of molality emerged from early 20th-century efforts in to quantify concentrations in ways that better suited thermodynamic analyses, particularly for like , lowering, , and . These properties depend primarily on the ratio of solute particles to molecules, making mass-based measures of more appropriate than volume-based ones, especially in non-aqueous systems or under varying temperatures where volumes change due to . In the early 1900s, prominent physical chemists such as advocated for solvent-mass-based concentrations to simplify calculations in these contexts, addressing the shortcomings of earlier measures like (equivalents per liter of ), which were prone to inconsistencies in mixed solvents or temperature fluctuations. , introduced in the late , relied on volume and reactive equivalents, limiting its utility for precise thermodynamic modeling of dilute s. and collaborators emphasized molal scales to normalize data across different solvents and conditions, facilitating comparisons in studies of solution ideality. Terms like "molal concentration" gained traction in subsequent decades for similar solvent-mass-based expressions in experimental work on electrolytes and nonelectrolytes. The noun "molality" was formally introduced by G. N. Lewis and Merle Randall in their seminal 1923 text Thermodynamics and the Free Energy of Chemical Substances, where it was defined as the number of moles of solute per kilogram of solvent to streamline free energy and equilibrium computations in solution thermodynamics. This innovation combined "mole" with the adjectival suffix "-al," analogous to "molarity" but tied to solvent mass rather than solution volume. The adoption of molality accelerated through the mid-20th century, with the International Union of Pure and Applied Chemistry (IUPAC) incorporating it into standardized terminology by the 1970s, building on earlier provisional uses to ensure consistency in reporting. This evolution underscored molality's role in bridging experimental observations with theoretical frameworks, particularly for colligative effects where solute-solvent interactions dominate.

Units and Notation

The SI unit for molality is moles of solute per kilogram of solvent (mol/), a derived unit within the (SI) that became officially recognized following the adoption of the as the seventh base unit at the 14th General Conference on Weights and Measures (CGPM) in 1971. This unit emphasizes mass-based measurement, aligning with the SI's foundational principles of using the for mass and the for . In standard notation, molality is denoted by the lowercase italic letter m, frequently subscripted to specify the solute, as in mi for the molality of the i-th component in a multicomponent . To prevent ambiguity with the symbol for mass (m), the International Union of Pure and Applied Chemistry (IUPAC) recommends the symbol b for molality, though m persists in widespread and practice. This notation clearly differentiates molality from molarity, which uses uppercase M or lowercase c for moles per liter of solution. Molality values are typically reported to three or four places to reflect the precision of analytical balances used in preparation, with the of the treated as the exact quantity in calculations. The "molal" describes or quantities based on molality, such as molal , in contrast to "" for volume-based measures like molarity. Unlike concentration units involving volume, molality excludes any volumetric dependence, ensuring consistency across temperature variations.

Practical Usage

Advantages Over Other Measures

One primary advantage of molality is its independence from variations, as it is defined in terms of the of rather than , which expands or contracts with thermal changes. This stability makes molality particularly suitable for calculations involving , such as and , where precise concentration measures are essential across a range of temperatures without requiring adjustments for shifts. Similarly, molality remains unaffected by pressure changes, since mass ratios do not alter under or , unlike volume-based measures that can vary in high-pressure environments or systems with dissolved gases. This property is beneficial in applications like geochemical studies or under elevated pressures, ensuring consistent concentration assessments. In multi-component , molalities exhibit additivity, meaning the total molality is simply the sum of individual solute molalities relative to the same mass, without needing corrections or recalculations for interactions. This simplifies mixture analysis, especially for non-ideal solutions like electrolytes, where molality provides a reliable basis for computations via osmolality, accounting for ion dissociation more accurately than volume-dependent units. Molality is preferred in situations involving concentrated solutions, such as some biochemical studies of proteins, where the volume of the solution may not be simply additive due to significant solute volume contributions, leading to variations in total solution density and imprecise volume measurements; using mass-based concentrations avoids these issues.

Limitations and Challenges

Preparing molal solutions demands precise measurement of the solvent's , which can be particularly difficult in practice. Volatile solvents, such as certain liquids, are prone to during weighing, leading to inaccuracies in the determined molality. Similarly, hygroscopic solvents absorb atmospheric , causing the measured to fluctuate and complicating the preparation process. These issues make molality less practical for routine laboratory work compared to volume-based measures like molarity. Although the total mass of a is fundamentally the sum of the solute and masses—ensuring additivity at the macroscopic level—solute-solvent interactions, such as or , can influence the effective behavior of the , indirectly affecting how molality is applied in non-ideal systems. This requires additional corrections for accurate thermodynamic calculations, adding complexity to its use. Molality proves less intuitive and practical for dilute , where the solute contribution to the total volume is negligible, and measuring volumes with pipettes or burettes is simpler and more straightforward than weighing small masses of . In such cases, molarity often suffices without the added effort of mass determinations. In multicomponent systems, determining the solvent mass for molality calculations presents significant analytical challenges, as isolating the primary from multiple components typically requires advanced separation techniques or spectroscopic methods to avoid errors in . Without pure substances to clearly define solute and , molality becomes ambiguous and difficult to apply accurately. In , fractions are often preferred for their simplicity in handling large-scale operations, ease of calculation without molecular weights, and direct compatibility with weighing-based . Molality is employed in specific applications, such as CO2 capture and textile dyeing, where its independence from and is beneficial. Additionally, some software prioritizes molarity or -based units over molality.

Relations to Other Concentration Quantities

To Mass Fraction

The mass fraction of a solute, denoted as w, is defined as the ratio of the mass of the solute to the total mass of the solution in a binary mixture./13%3A_Solutions/13.03%3A_Units_of_Concentration) To derive the relationship between molality and mass fraction, consider a binary solution containing 1 kg (1000 g) of solvent and m moles of solute, where m is the molality. The mass of the solute is then m \times M, with M being the molar mass of the solute in g/mol. The total mass of the solution is $1000 + m \times M g, yielding the conversion formula: w = \frac{m M}{1000 + m M} This formula links molality directly to mass fraction and originates from standard concentration relations in chemical engineering references. The inverse relation, solving for molality in terms of mass fraction, is: m = \frac{1000 w}{M (1 - w)} These conversions assume ideal mixing with no volume change upon dissolution and apply specifically to binary solutions consisting of one solute and one solvent. As an illustrative example, consider a 1 molal aqueous solution of sodium chloride (NaCl), where the molar mass M of NaCl is 58.44 g/mol. Substituting into the formula gives w = \frac{1 \times 58.44}{1000 + 1 \times 58.44} = \frac{58.44}{1058.44} \approx 0.055.

To Mole Fraction

The mole fraction x of a solute in a solution represents the proportion of moles of the solute to the total moles of all components, defined as x = \frac{n_\text{solute}}{n_\text{total}}, where n_\text{total} = n_\text{solute} + n_\text{solvent} + \sum n_\text{other}} for multicomponent systems, though it simplifies to x = \frac{n_2}{n_1 + n_2} for binary solutions with solvent (1) and solute (2). This mole-based measure provides a dimensionless quantity that is particularly valuable for expressing compositions in thermodynamic analyses, as it directly reflects the relative numbers of particles independent of molecular size. In a binary solution, the mole fraction of the solute can be expressed in terms of molality m (moles of solute per kg of solvent) using the relation x = \frac{m \cdot M_\text{solvent}}{1000 + m \cdot M_\text{solvent}}, where M_\text{solvent} is the of the in g/mol, and the denominator accounts for the total moles assuming 1 kg (1000 g) of , yielding n_\text{solvent} = 1000 / M_\text{solvent}. This formula derives from substituting the definitions: n_2 = m (for 1 kg ) and n_1 = 1000 / M_\text{solvent}, directly linking the mass-independent molality to the additive mole proportions. For dilute solutions, where m is small such that m \cdot M_\text{solvent} \ll 1000, the formula approximates to x \approx \frac{m \cdot M_\text{solvent}}{1000}, simplifying calculations by neglecting the solute's contribution to the total moles./16%3A_Aqueous_Equilibria/16.09%3A_Molality_and_Mole_Fraction) This approximation is common in introductory contexts. The relation between molality and is especially useful in , facilitating derivations like , where the partial of the solvent is P = x_\text{solvent} \cdot P^\circ, with x_\text{solvent} = 1 - x for systems, allowing molality-based to predictions./Physical_Properties_of_Matter/Solutions_and_Mixtures/Ideal_Solutions/Changes_In_Vapor_Pressure_Raoult%27s_Law) This exact relation holds for binary solutions; in multicomponent systems, the mole fraction expression extends to include additional components but requires more complex derivations, as covered in advanced applications.

To Molarity

Molarity, denoted as c, is defined as the number of moles of solute (n_\text{solute}) divided by the volume of the solution in liters (V_\text{solution}), expressed as c = \frac{n_\text{solute}}{V_\text{solution}}. The relationship between molality m and molarity c arises from the dependence of solution volume on its total mass and density. For a solution with molality m (moles of solute per kilogram of solvent), the total mass per kilogram of solvent is $1000 + m M_\text{s} grams, where M_\text{s} is the molar mass of the solute in g/mol. The volume in liters is then this mass divided by the solution density \rho (in g/mL) and scaled appropriately, yielding the formula c = \frac{m \rho}{1 + \frac{m M_\text{s}}{1000}}. This equation highlights how molarity incorporates the solution's , which accounts for the volume contribution of both solute and . Unlike molality, which remains because it relies on fixed masses of solute and , molarity varies with due to changes in solution via and shifts. In aqueous solutions, typically decreases as rises, expanding and increasing molarity for a given molality. For instance, in NaCl(aq) systems, experimental data show that at molality drops from about 1.038 g/mL at 20°C to 1.025 g/mL at 40°C, resulting in a corresponding rise in molarity. Molality is preferred for precise colligative property calculations, such as or lowering, as these effects depend directly on the mole ratio in the mass, unaffected by volume fluctuations. Conversely, molarity is standard for titrations, where reagent volumes are measured directly to determine reaction . Molality has gained favor for non-aqueous s, where densities exhibit greater sensitivity to and solute addition compared to , making volume-based measures like molarity less reliable.

To Mass Concentration

Mass concentration, denoted as \rho_{\text{solute}}, represents the mass of solute per unit volume of the solution, calculated as \rho_{\text{solute}} = \frac{m_{\text{solute}}}{V_{\text{solution}}}, where m_{\text{solute}} is the mass of the solute and V_{\text{solution}} is the volume of the solution; it is typically expressed in units such as grams per liter (g/L) or milligrams per liter (mg/L). This measure is particularly useful in contexts requiring direct assessment of solute mass dispersed in a given volume, such as regulatory monitoring of contaminants. To relate molality (m) to mass concentration, the (\rho) of the and the (M_{\text{solute}}) of the solute are required, as depends on the and . For a with molality m (moles of solute per kilogram of ), the mass concentration is given by: \rho_{\text{solute}} = \frac{m \cdot \rho \cdot M_{\text{solute}}}{1000 + m \cdot M_{\text{solute}}} where \rho is the of the in g/mL, M_{\text{solute}} is in g/mol, and the denominator accounts for the of 1 kg plus solute (with 1000 g/kg). This derives from expressing the solute as m \cdot M_{\text{solute}} g per kg , as $1000 + m \cdot M_{\text{solute}} g, and as divided by . In dilute solutions, where m \cdot M_{\text{solute}} \ll 1000, the formula simplifies to \rho_{\text{solute}} \approx \frac{[m](/page/M) \cdot \rho \cdot M_{\text{solute}}}{1000}, assuming the density approximates that of the (often 1 g/mL for water-based systems). This approximation facilitates quick estimates but loses accuracy at higher concentrations where solute contributions to volume and become significant. In , mass concentration serves as the standard metric for tracking pollutants, such as fine (PM_{2.5}) in air or trace metals in , enabling compliance with regulations like those set by the U.S. Environmental Protection Agency, which specify limits in \mug/m³ or mg/L. For instance, seasonal variations in PM_{2.5} mass concentrations are monitored to assess sources and impacts in areas. A key distinction is that molality circumvents the need for measurements, relying instead on mass, which simplifies calculations in scenarios where volume data is imprecise or temperature-variable.

Conversion Formulas and Examples

To convert molarity to molality for a specific , the of the and the of the solute are essential. Consider a 2.0 M (HCl) with a of 1.03 g/mL at 20°C and of HCl = 36.46 g/mol. The step-by-step calculation is as follows:
  1. Determine the mass of 1 L of :
    Mass = × = 1.03 g/mL × 1000 mL = 1030 g.
  2. Calculate the mass of solute in 1 L:
    of HCl = molarity × = 2.0 mol/L × 36.46 g/mol = 72.92 g.
  3. Calculate the of ():
    of = total - solute = 1030 g - 72.92 g = 957.08 g = 0.95708 kg.
  4. Calculate molality:
    m = \frac{\text{moles of solute}}{\text{kg of solvent}} = \frac{2.0 \ \text{[mol](/page/Mol)}}{0.95708 \ \text{kg}} \approx 2.09 \ m
This example demonstrates how molarity slightly underestimates molality for this concentration due to the solution's exceeding 1 g/mL. In ideal dilute aqueous solutions at 25°C, where the density is approximately 1 g/mL and the solute contribution to volume is negligible, molality (m) equals molarity (c). This equality holds because 1 L of solution ≈ 1 kg of solvent. The table below illustrates this for select dilute concentrations of a nonelectrolyte like glucose (molar mass 180.16 g/mol) in water, assuming ideal behavior (density ≈ 1.00 g/mL).
Molarity (c, mol/L)Molality (m, mol/kg)Ratio (m/c)
0.0010.0011.00
0.010.011.00
0.10.11.00
0.50.51.00
These values are approximate for concentrations below 0.5 M, where deviations remain under 0.5%./11%3A_Liquids_Solids_and_Intermolecular_Forces/11.06%3A_Molality_Molarity_and_Mole_Fraction) For multi-step conversions, such as from molality to using mass fraction as an intermediate, consider a 1.0 m solution of (NaCl, 58.44 g/mol) in ( 18.02 g/mol). Assume the density is not needed for this path, as it relies on mass ratios.
  1. Calculate the mass fraction of solute (w_solute):
    w = \frac{m \times M_\text{solute}}{1000 + m \times M_\text{solute}} = \frac{1.0 \times 58.44}{1000 + 1.0 \times 58.44} = \frac{58.44}{1058.44} \approx 0.0552
  2. Mass fraction of solvent (w_solvent) = 1 - w_solute = 0.9448.
  3. Moles of solute per gram of solution = w_solute / M_solute = 0.0552 / 58.44 ≈ 0.000944 /g.
  4. Moles of solvent per gram of solution = w_solvent / M_solvent = 0.9448 / 18.02 ≈ 0.05243 /g.
  5. Mole fraction of solute (x_solute):
    x = \frac{\text{moles solute per g}}{\text{moles solute per g} + \text{moles solvent per g}} = \frac{0.000944}{0.000944 + 0.05243} \approx 0.0177
This method is useful when direct mole counts are cumbersome for complex systems. Conversions typically require molar masses from standard references like the CRC Handbook and solution densities from experimental data or tables (e.g., density references). For complex cases involving multiple solutes or temperature variations, software tools such as or Python scripts with libraries like can automate calculations, inputting variables for iterative error checking. Error propagation in these conversions is critical, particularly from density measurements, which often have uncertainties of ±0.001 g/mL. For the molarity-to-molality formula m = \frac{c}{\rho - c \cdot (M/1000)}, the relative uncertainty in molality is approximately \frac{\delta m}{m} \approx \frac{\delta c}{c} + \frac{\delta \rho \cdot \rho}{\rho - c \cdot (M/1000)} \cdot \frac{1}{\rho}, derived from partial derivatives. In the 2.0 M HCl example above, a 0.001 g/mL uncertainty in density propagates to about ±0.02 m in molality (≈1% relative error), highlighting the sensitivity in non-dilute solutions./Quantifying_Nature/Significant_Digits/Propagation_of_Error)

To Osmolality

Osmolality represents an extension of molality that accounts for the effective number of solute particles contributing to , such as , in a . Defined as the number of osmoles of solute per of , osmolality is calculated using the : \text{osmolality} = m \times \nu where m is the molality in moles per of , and \nu (the van't Hoff ) is the average number of particles (ions or molecules) into which each of solute dissociates in . For non-electrolytes like glucose, \nu = 1, so osmolality equals molality; for electrolytes like (NaCl), which dissociates into two s, \nu \approx 2, though the exact value may deviate slightly from ideality due to ion pairing at higher concentrations. This distinction highlights a key difference: molality measures the concentration based on formula units of solute, while osmolality reflects the total particle count influencing osmotic behavior. In physiological contexts, osmolality is crucial for maintaining across membranes, with normal human blood osmolality ranging from 275 to 295 mOsm/kg to prevent excessive movement. Deviations can lead to conditions like or , disrupting cellular function. In , osmolality guides the formulation of intravenous (IV) solutions to match plasma levels and avoid or ; for instance, a 0.9% NaCl (normal saline) solution has an osmolality of approximately 286 mOsm/kg, closely approximating physiological conditions despite its theoretical value of 308 mOsm/kg based on ideal dissociation (m \approx 0.154 mol/kg, \nu = 2). This adjustment accounts for non-ideal behavior in aqueous solutions. Osmolality is commonly measured using freezing point depression osmometry, which exploits the colligative property where the freezing point of a lowers proportionally to the solute particle concentration. The relationship is expressed as: \Delta T_f = K_f \times \text{osmolality} where \Delta T_f is the and K_f is the 's (1.86 °C kg/osmol for ). This method provides a direct, non-specific assessment of total osmotic activity, essential for clinical diagnostics and in biological and pharmaceutical samples.

Advanced Applications

Apparent Molar Properties

In solution thermodynamics, the apparent molar volume V_\phi quantifies the contribution of a solute to the total volume of a beyond that expected from the pure , facilitating analysis of volumetric non-idealities when concentrations are expressed in molal units. It is defined as V_\phi = \frac{V - n_{\text{solvent}} V_{\text{solvent}}^*}{n_{\text{solute}}} where V is the total volume of the solution, n_{\text{solvent}} and n_{\text{solute}} are the moles of and solute, respectively, and V_{\text{solvent}}^* is the of the pure . When using molality m (moles of solute per of ), this expression is conveniently scaled for 1 kg of , yielding n_{\text{solute}} = m and n_{\text{solvent}} \approx 55.51 mol for , allowing V_\phi to reflect solute-solvent interactions per molal unit without volume contraction effects from changing solvent density. The partial molar volume V_m of the solute is related to the apparent molar volume by V_m = V_\phi + n_{\text{solute}} \left( \frac{\partial V_\phi}{\partial n_{\text{solute}}} \right)_{n_{\text{solvent}}, T, P}. At infinite dilution, V_\phi = V_m, but they deviate at finite concentrations due to solute- interactions. Experimentally, V_\phi is determined from precise measurements of solutions across a range of molalities, using \rho = \frac{\text{total mass}}{V} to compute V, followed by substitution into the defining equation; pycnometric or vibrating-tube densimetry provides the required accuracy for systems. This approach is particularly valuable for predicting non-ideal behavior in solutions, where V_\phi deviations from additivity reveal shells, pairing, and electrostrictive effects that alter solution . A representative example is aqueous NaCl solutions, where V_\phi increases with molality due to progressive of ions and diminished volume contraction from overlapping hydration layers; at low m (e.g., 0.1 mol kg⁻¹), V_\phi \approx 16.6 cm³ mol⁻¹, significantly below the solid NaCl of 27 cm³ mol⁻¹, but rises roughly linearly with \sqrt{m} up to , indicating weakening non-ideal contributions at higher concentrations. Analogously, the apparent molar enthalpy H_\phi extends this framework to energetic properties, defined as H_\phi = \frac{H - n_{\text{solvent}} H_{\text{solvent}}^*}{n_{\text{solute}}} where H is the total of the solution and H_{\text{solvent}}^* is the of the pure ; in molal terms, it is derived from calorimetric measurements of heats of dilution or solution heat capacities at varying m, enabling dissection of enthalpic non-idealities from and in electrolytes. Like V_\phi, H_\phi is crucial for modeling non-ideal , as its concentration dependence informs exothermic or endothermic mixing processes in .

Activity Coefficients in Molal Terms

In non-ideal solutions, the molal activity coefficient, denoted as \gamma_{m,i} for solute i, accounts for deviations from ideal behavior by modifying the chemical potential according to the relation \mu_i = \mu_i^* + RT \ln (m_i \gamma_{m,i}), where \mu_i^* is the standard chemical potential, R is the , T is the , and m_i is the molality of the solute. This formulation ensures that the activity a_i = m_i \gamma_{m,i} replaces molality in thermodynamic expressions for non-ideal systems, maintaining consistency with the molal scale's independence from solvent . The molal activity coefficient differs from the mole fraction-based coefficient \gamma_{x,i} particularly in dilute solutions, where \gamma_{m,i} \approx \gamma_{x,i} (1 - x_{\text{solute}}) and x_{\text{solute}} is the solute mole fraction, reflecting the scale's focus on solvent mass rather than total composition. For electrolyte solutions, the Debye-Hückel limiting law adapted to the molal scale predicts mean activity coefficients via \log \gamma_{\pm} = -A |z_+ z_-| \sqrt{m}, where \gamma_{\pm} is the mean ionic activity coefficient, A is a - and solvent-dependent constant (approximately 0.509 for at 25°C), z_+ and z_- are charges, and m is the total molality. This law applies effectively at low concentrations, typically below 0.01 molal, providing a theoretical basis for estimating non-ideal effects in aqueous systems. In practical applications, molal activity coefficients enable the expression of equilibrium constants in molal units, such as for solubility products where K_{sp} = \prod (m_i \gamma_{m,i})^{ \nu_i } for the dissolution reaction, facilitating accurate predictions of sparingly soluble salt behaviors without density corrections. For instance, tabulated values for in show \gamma_{m,\pm} decreasing from near 1 at infinite dilution to about 0.657 at 1 molal and 0.421 at 5 molal at 25°C, illustrating the increasing non-ideality with concentration. Similar data for other salts like indicate \gamma_{m,\pm} \approx 0.604 at 1 molal, underscoring the role of ion-specific interactions in these coefficients.

Multicomponent Solutions

In multicomponent solutions, molality is extended from systems to describe the concentration of multiple solutes dissolved in a single . Each solute i has an individual molality m_i, defined as the number of moles of solute i (n_i) per of , analogous to the binary case but applied independently to each component. The total molality m_{\text{total}} of a multicomponent is the sum of the individual molalities of all solutes: m_{\text{total}} = \sum_i m_i = \sum_i \frac{n_i}{m_{\text{solvent}}} where m_{\text{solvent}} is the mass of the solvent in kilograms. This total provides an overall measure of solute concentration, useful for comparing solution strengths across systems with varying solute compositions. For instance, in electrolyte mixtures, total molality influences colligative properties like osmotic pressure. The solvent in multicomponent systems is typically defined as the major component by mass, in which the solutes are dissolved, ensuring the solution behaves as a dilute or semi-dilute mixture relative to that phase. However, challenges arise in mixed-solvent systems, where multiple components could qualify as solvents, such as in co-solvent mixtures or azeotropes; here, precise identification of the solvent mass requires experimental determination of phase behavior to avoid ambiguities in molality calculations. In thermodynamic analyses of multicomponent solutions, the contribution of a specific solute can be isolated by varying its molality m_i while keeping the molalities of other solutes m_j (for j \neq i) constant. Molality finds practical applications in analyzing multicomponent natural and engineered systems. In , which contains multiple salts like NaCl, MgSO₄, and CaCl₂ dissolved in , total molality is calculated from measurements to model osmotic behavior and ion transport; for example, standard seawater has a total molality of approximately 1.0 mol/kg, aiding in oceanographic studies. Similarly, in electrolytes, such as lithium-ion systems with salts like LiPF₆ in solvents, molality quantifies high-concentration effects (often >10 mol/kg) on and , where distinctions from molarity are critical for performance optimization. For non-ideal multicomponent solutions, especially electrolytes, deviations from ideality are accounted for using mean activity coefficients (\gamma_\pm), which average the activity coefficients of ions in the on a molal scale. These coefficients, derived from measurements of osmotic coefficients or , correct for interionic interactions; for uni-univalent salts at 25°C, \gamma_\pm decreases with increasing total molality, as seen in NIST compilations for aqueous NaCl up to 6 mol/kg.

Binary to Multicomponent Derivations

The extension of the Pitzer ion-interaction model from binary to multicomponent electrolyte solutions in molal terms relies on incorporating pairwise (binary) and triplet (ternary) interaction parameters into the expression for the excess Gibbs energy, allowing predictions of thermodynamic properties like activity and osmotic coefficients for mixtures. In this framework, the excess Gibbs energy is formulated in molal units as \frac{G^\text{ex}}{RT m_\text{solv}} = f(\sum m_i, \beta_{ij}(I), C_{ijk}, \theta_{ij}), where m_\text{solv} is the molality of the solvent (typically 55.51 mol·kg⁻¹ for water), m_i are the molalities of ionic species, \beta_{ij}(I) are ionic-strength-dependent binary interaction parameters, C_{ijk} are ternary parameters, and \theta_{ij} are mixing terms for ions of the same charge sign. This form generalizes the binary case, where only self-interactions within a single electrolyte are considered, by summing contributions over all species and adding cross-interaction terms derived or estimated from binary subsystem data. To derive multicomponent molalities or related properties like apparent molal volumes or osmotic coefficients from binary data, approximations often assume initial additivity of binary contributions, adjusted by empirical cross coefficients. For instance, in ternary systems, the effective pairwise interaction molality m_{ij}^\text{ternary} is approximated as m_{ij}^\text{ternary} \approx m_i^\text{binary} + \Delta m_{ij}, where \Delta m_{ij} incorporates correction terms from binary-derived cross coefficients such as \theta_{ij} or \psi_{ijk} (for ternary mixing). These corrections are obtained by fitting binary osmotic or data to the and then extrapolating to mixtures, often using constant or equivalents as scaling variables to minimize deviations. A key equation for apparent molar properties in multicomponent systems, such as the apparent osmotic \phi^\text{ternary}, takes the form \phi^\text{ternary} = \sum_i x_i \phi_i^\text{binary} + \sum_{i < j} x_i x_j \delta_{ij}, where x_i are mole fractions of components, \phi_i^\text{binary} is the apparent property from the pure subsystem at equivalent concentration, and \delta_{ij} are mixing corrections (e.g., derived from \theta_{ij} parameters). This derivation assumes weak non-additive effects, enabling prediction of mixture without full ternary experimental data, though \delta_{ij} must be calibrated from limited mixture measurements or theoretical estimates. An illustrative example is the prediction of molalities and osmotic coefficients in the NaCl-KCl-H₂O system using Pitzer parameters. for NaCl-H₂O and KCl-H₂O provide \beta_{NaCl}, \beta_{KCl}, and self-interaction terms, which are combined with the cross coefficient \theta_{NaK} \approx 0.012 (fitted from isopiestic measurements) to account for Na⁺-K⁺ repulsions; ion-pair parameters like those for NaCl⁰ or KCl⁰ are negligible here but included if occurs at high concentrations. This approach accurately predicts mixture osmotic coefficients within 1-2% up to total molalities of 6 mol·kg⁻¹ at 25°C, as validated against experimental data. These derivations assume approximate additivity of interactions, which holds for dilute to moderate concentrations in simple mixtures but fails in systems with strong specific interactions, such as formation (e.g., in sulfate-phosphate mixtures where \psi terms dominate and require direct measurement). In such cases, the model overpredicts solubilities or underestimates activity coefficients by up to 10-20% without additional adjustments.

References

  1. [1]
    molality (M03970) - IUPAC Gold Book
    Amount of entities of a solute divided by the mass of the solvent. Sources: Green Book, 2 nd ed., p. 42
  2. [2]
    Colligative Properties – Chemistry - UH Pressbooks
    ... A total mol of all components. Molality is a concentration unit defined as the ratio of the numbers of moles of solute to the mass of the solvent in kilograms:.Mole Fraction And Molality · Vapor Pressure Lowering · Osmosis And Osmotic Pressure...
  3. [3]
    13.5 Colligative Properties of Solutions
    Second, molality and mole fraction are proportional for relatively dilute solutions, but molality has a larger numerical value (a mole fraction can be only ...Vapor Pressure Of Solutions... · Freezing Point Depression · Osmotic Pressure
  4. [4]
    16.11: Molality - Chemistry LibreTexts
    Mar 20, 2025 · The molality ( m ) of a solution is the moles of solute divided by the kilograms of solvent. · The value of molality does not change with ...
  5. [5]
    Molality Explained: Definition, Examples, Practice & Video Lessons
    May 11, 2022 · Molality is a measure of solution concentration, defined as the number of moles of solute per kilogram of solvent, making it temperature-independent since it ...
  6. [6]
    The Chemical Activity of the Ions of Hydrochloric Acid ... - jstor
    at the smallest concentration (0.00167 molal) the activity-coefficient is equal to the ionization-coefficient (0.988) derived from the ratio of the ...
  7. [7]
    manual of symbols and terminology - iupac
    Apr 2, 2008 · 1963—1969 W. .Jaenicke ... (3) See Section 1.2. (4) A solution having a molality equal to 0.1 mol kg' is sometimes called a 0.1 molal.
  8. [8]
    Resolution 3 of the 14th CGPM (1971) - BIPM
    SI unit of amount of substance (mole)​​ The mole is the amount of substance of a system which contains as many elementary entities as there are atoms in 0.012 ...Missing: molality per
  9. [9]
    [PDF] Guide for the Use of the International System of Units (SI)
    Feb 3, 1975 · SI unit: kilogram per mole (kg/mol). Definition: mass of a substance ... A.7 Mole (14th CGPM, 1971). 1. The mole is the amount of ...
  10. [10]
    [PDF] Quantities, Units and Symbols in Physical Chemistry - IUPAC
    (14) In this definition the symbol m is used with two different meanings: mB denotes the molality of solute B, mA denotes the mass of solvent A (thus the ...
  11. [11]
    Advantages and Disadvantages of Using Molality - BYJU'S
    The use of this property was first recorded in the publication authored by G. N. Lewis and M. Randall in the year 1923. The topic was covered in the book ...
  12. [12]
    Molality - an overview | ScienceDirect Topics
    Molality (m) is defined as the number of moles of solute per kilogram of solvent, serving as a concentration unit that remains independent of temperature ...Missing: history | Show results with:history
  13. [13]
    Why do we use molality on describing colligative property?
    Sep 19, 2020 · Molality is a useful unit in colligative properties because it is independent of temperature, it is not affected by volume changes upon mixing.
  14. [14]
    Osmolality and Osmolarity - BYJU'S
    The term “osmolality” is generally preferred because the osmotic pressure of a solution is considerably more closely related to its osmolality than to its ...
  15. [15]
    Why Is Molality Used Instead of Molarity? - ThoughtCo
    Dec 7, 2019 · Molality is preferred when solutions are concentrated, use non-water solvents, or have temperature changes, as the mass of solute and solvent ...Missing: protein | Show results with:protein<|control11|><|separator|>
  16. [16]
    Molality: Definition, Formula, and Solved Problems
    Jan 25, 2025 · Molality has several disadvantages that limit its practicality. One major drawback is its dependence on precise measurements of solvent mass. ...
  17. [17]
    Molality: Definition, Formula, Advantages and Mole Fraction
    The first time this attribute is mentioned in a publication is in a 1923 work by G. N. Lewis and M. Randall. The topic was covered in the book ...
  18. [18]
    Perry's Chemical Engineers' Handbook | McGraw-Hill Education
    This industry-standard resource, first published in 1934, has equipped generations of engineers and chemists with vital information, data, and insights.Missing: molality | Show results with:molality
  19. [19]
    Sodium Chloride | ClNa | CID 5234 - PubChem
    Sodium chloride, for molecular biology, DNase, RNase, and protease, none ... Exact Mass. Property Value. 57.9586220 Da. Reference. Computed by PubChem 2.2 ...
  20. [20]
    Concentrations of Solutions
    The mass of solvent (in kilograms) in the solution. To calculate molality we use the equation: Equation for calculating molality. Top. Mole Fraction. The mole ...
  21. [21]
    [PDF] Chapter 25 solutions
    A/% p/g-mL c/mol·L-1 25-8. Derive a relation between the mole fraction of the solvent and the molality of a solution. V/mL = 1001.70 + (17.298 kg-mol¨¹)m + (0. ...
  22. [22]
    Solutions Help Page
    Molality, m, tells us the number of moles of solute dissolved in exactly one kilogram of solvent. (Note that molality is spelled with two "l"'s and represented ...<|control11|><|separator|>
  23. [23]
    [PDF] Thermodynamic Properties of Aqueous Sodium Chloride Solutions
    Oct 15, 2009 · ... Temperature DependenCe of Parameters and Weighting of Data ... solution of NaCl(aq) ................... 65. A-12. Relative entropy of ...
  24. [24]
    [PDF] 14 Solutions
    We simply use kilograms of solvent instead of liters of solvent. We introduce molality here because in the study of colligative properties we use it in ...
  25. [25]
    Acid-Base Titration 1
    Calculate the molarity of an acetic acid solution if 34.57 mL of this solution are needed to neutralize 25.19 mL of 0.1025 M sodium hydroxide.<|control11|><|separator|>
  26. [26]
    [PDF] QSEARCH - Emporia ESIRC - Emporia State University
    environment of the ion one type of solvent molecule is preferred. ... in non-aqueous or mixed solvents. In these ... weight, since molality is temperature ...
  27. [27]
    NIST Guide to the SI, Chapter 8
    Jan 28, 2016 · Quantity symbol:M. SI unit: kilogram per mole (kg/mol). Definition: mass of a substance divided by its amount of substance: M = m/ ...
  28. [28]
    11. Chemical concentrations — Mathematics for Natural Sciences B
    The concentration of a solution may also be given as a mass fraction %(w/w), which is the solute mass as a percentage of the total solution mass. To convert ...
  29. [29]
    Spatial and seasonal variability of the mass concentration and ...
    The seasonal changes in ambient mass concentrations and chemical composition of fine particulate matter (PM2.5) were investigated in three locations in Poland.
  30. [30]
    [PDF] Investigating the PM2.5 mass concentration growth processes ...
    Jan 4, 2019 · 5 mass concentration growth processes in the NCP and YRD regions. The results will improve our understanding of the mechanism of severe haze ...
  31. [31]
  32. [32]
    Calculations involving molality, molarity, density, mass percent, mole ...
    Calculations involving molality, molarity, density, mass percent, mole fraction. Problems #1 - 10 · 1) Determine volume of thiophene: 9.660 g ÷ 1.065 g/mL = 9.07 ...
  33. [33]
  34. [34]
    Hydrochloric acid density - Anton Paar Wiki
    Density of hydrochloric acid – concentration measurement ; 1.0105, 2.5 ; 1.0130, 3.0 ; 1.0154, 3.5 ; 1.0179, 4.0.
  35. [35]
    Comparison of vapour pressure osmometry, freezing point ...
    Oct 1, 2022 · The relationship between osmolality and van't Hoff factor is show in Eq. (6) where Osm/kg is osmolality, i* is the van't Hoff factor from ...
  36. [36]
    Non-ideal Solution Thermodynamics of Cytoplasm - PMC
    ... van't Hoff factor that accounts for dissociation). In an attempt to allow ... where m(π) is the molality as a function of osmolality and b* is the ...
  37. [37]
    Serum Osmolality - StatPearls - NCBI Bookshelf - NIH
    Feb 27, 2024 · [1] The term osmolality expresses concentrations relative to the mass of the solvent, whereas the term osmolarity expresses concentrations per ...
  38. [38]
    Intravenous fluids: issues warranting concern - PMC - NIH
    0.9% saline has a theoretical osmolarity of 308 mosmol l−1 (=154 mmol l−1 Na+ + 154 mmol l−1 Cl−) but an actual osmolality of 286 mosmol kg−1 H2O−1.
  39. [39]
    Freezing Point Depression Theory - Advanced instruments
    By precisely measuring the freezing point of the solution, the osmolality, or concentration, can be determined. Freezing point measurement is non-specific; that ...
  40. [40]
    [PDF] Volumetric Properties of Aqueous Sodium Chloride Solutions
    Molality, mol kg". 1. Avogadro's number. Moles of solvent, solute ... The apparent molal volume is defined as. V n 1 1. R2 so that from eq (9),. 1 ag.
  41. [41]
    [PDF] Partial Molar Volume - Chemistry at URI
    molar volume is important, and define partial and apparent molar volume and explain the relationship between them. • Signed procedure and signed data.
  42. [42]
    Investigations of the (p, ρ, T) Properties and Apparent Molar ...
    The (p, ρ, T) properties and apparent molar volumes Vϕ of LiCl in ethanol at T = (298.15 to 398.15) K and pressures up to p = 40 MPa are reported.
  43. [43]
    CHAPTER 18: Molar Volumes of Electrolyte Solutions - Books
    Second, the molar volumes of electrolyte solutions provide valuable insights into ion–solvent interactions (at infinite dilution), ion–ion interactions (at ...
  44. [44]
    The apparent molar heat capacities and volumes of aqueous NaCl ...
    The heat capacities per unit volume of aqueous solutions of NaCl were measured with a flow microcalorimeter. The molality and temperature range covered were ...
  45. [45]
    Thermodynamic properties of multicomponent NaCl–LiCl–H2O ...
    The relative humidities of this system are measured at the total molalities from 0.3 mol kg−1 to about saturation for different ionic-strength fractions y of ...
  46. [46]
    The effect of pressure on transition metals in seawater - Academia.edu
    The partial molality of the added salt (m3) are tabulated in Tables 1 and 2. molal volumes and compressibility measurements in water have been made for a ...
  47. [47]
    gsw_molality_from_SA - TEOS-10
    DESCRIPTION: Calculates the molality of seawater. TEOS-10, Click for a more detailed description of the molality of seawater ...
  48. [48]
    [PDF] Osmotic Coefficients and Mean Activity Coefficients of Uni-univalent
    This paper gives values for the osmotic coefficients and mean activity coefficents of uni-univalent electrolytes in aqueous solutions at 25 °C. The values are ...
  49. [49]
    Chemical speciation models based upon the Pitzer activity ...
    Jul 20, 2022 · Pitzer-based models of multicomponent electrolyte solutions are complex, and involve many interactions and parameters. The models of ...
  50. [50]
    Evaluation of Pitzer ion interaction parameters of aqueous mixed ...
    Evaluation of Pitzer ion interaction parameters of aqueous mixed electrolyte solutions at 25.degree.C. 2. Ternary mixing parameters.
  51. [51]
    [PDF] Activity Coefficient Derivatives of Ternary Systems Based on ...
    May 7, 2007 · Fit the ternary data for example at constant molality using the constant molality binary approximation up to a total molality just below the ...
  52. [52]
    Comparison among Pitzer-type Models for the Osmotic and Activity ...
    In this work three Pitzer-type models were compared: the original one, the Archer model (an extended Pitzer model) with ionic strength dependence in the ...
  53. [53]
    [PDF] Aqueous electrolyte solution modelling: Limitations of the Pitzer ...
    Some very large Pitzer models Page 4 4 have been developed for multicomponent solutions, including one for seawater containing 16 major ionic components and a ...