12
7 REVIEW The assessment of acid-base analysis: comparison of the “traditional” and the “modern” approaches Jasna Todorović 1 , Jelena Nešovic-Ostojić 1 , Aleksandar Milovanović 2 , Predrag Brkić 3 , Mihailo Ille 4 , Dušan Čemerikić 1 1 Department of Pathological Physiology, 2 Insitute of Occupational Health, 3 Department of Medical Physiology, 4 Department of Surgery; School of Medicine, University of Belgrade, Belgrade, Serbia Corresponding author: Jelena Nešović Ostojić Department of Pathological Physiology, School of Medicine Dr. Subotića 1/II, 11000 Belgrade, Serbia Phone: +381 11 2685 282; Fax: +381 11 2685 340; E mail: [email protected] Original submission: 03 September 2014; Revised submission: 21 November 2014; Accepted: 18 December 2014. Med Glas (Zenica) 2015; 12(1):7-18 ABSTRACT Three distinct approaches are currently used in assessing acid-base disorders: the traditional - physiological or bicarbonate-centered approach, the base-excess approach, and the “modern” physico- chemical approach proposed by Peter Stewart, which uses the strong ion difference (particularly the sodium chloride difference) and the concentration of nonvolatile weak acids (particularly al- bumin) and partial pressure of carbon dioxide (pCO 2 ) as indepen- dent variables in the assessment of acid-base status. The traditio- nal approach developed from the pioneering work of Henderson and Hasselbalch and the base-excess are still most widely used in clinical practice, even though there are a number of problems identified with this approach. The approach works well clinically and is recommended for use whenever serum total protein, albu- min and phosphate concentrations are normal. Although Stewart’s approach has been largely ignored by physiologists, it is increasin- gly used by anesthesiologists and intensive care specialists, and is recommended for use whenever serum’s total protein, albumin or phosphate concentrations are markedly abnormal, as in critically ill patients. Although different in their concepts, the traditional and modern approaches can be seen as complementary, giving in prin- ciple, the same information about the acid-base status. Key words: acid-base balance, anion gap, strong ion difference, bicarbonate, base excess, nonvolatile weak acids, strong ion gap

02 Todorovic.pdf

Embed Size (px)

Citation preview

  • 7REVIEW

    The assessment of acid-base analysis: comparison of the traditional and the modern approachesJasna Todorovi1, Jelena Neovic-Ostoji1, Aleksandar Milovanovi2, Predrag Brki3, Mihailo Ille4, Duan emeriki1

    1Department of Pathological Physiology, 2Insitute of Occupational Health, 3Department of Medical Physiology, 4Department of Surgery;

    School of Medicine, University of Belgrade, Belgrade, Serbia

    Corresponding author:

    Jelena Neovi Ostoji

    Department of Pathological Physiology,

    School of Medicine

    Dr. Subotia 1/II, 11000 Belgrade, Serbia

    Phone: +381 11 2685 282;

    Fax: +381 11 2685 340;

    E mail: [email protected]

    Original submission:

    03 September 2014;

    Revised submission:

    21 November 2014;

    Accepted:

    18 December 2014.

    Med Glas (Zenica) 2015; 12(1):7-18

    ABSTRACT

    Three distinct approaches are currently used in assessing acid-base disorders: the traditional - physiological or bicarbonate-centered approach, the base-excess approach, and the modern physico-chemical approach proposed by Peter Stewart, which uses the strong ion difference (particularly the sodium chloride difference) and the concentration of nonvolatile weak acids (particularly al-bumin) and partial pressure of carbon dioxide (pCO2) as indepen-dent variables in the assessment of acid-base status. The traditio-nal approach developed from the pioneering work of Henderson and Hasselbalch and the base-excess are still most widely used in clinical practice, even though there are a number of problems identified with this approach. The approach works well clinically and is recommended for use whenever serum total protein, albu-min and phosphate concentrations are normal. Although Stewarts approach has been largely ignored by physiologists, it is increasin-gly used by anesthesiologists and intensive care specialists, and is recommended for use whenever serums total protein, albumin or phosphate concentrations are markedly abnormal, as in critically ill patients. Although different in their concepts, the traditional and modern approaches can be seen as complementary, giving in prin-ciple, the same information about the acid-base status.

    Key words: acid-base balance, anion gap, strong ion difference, bicarbonate, base excess, nonvolatile weak acids, strong ion gap

  • Medicinski Glasnik, Volume 12, Number 1, February 2015

    8

    INTRODUCTION

    Historically, two different conceptual approa-ches have been evolved among clinicians and physiologists for interpreting acid-base pheno-mena. The traditional or bicarbonate-centered approach relies qualitatively on the Henderson-Hasselbalch equation, whereas the modern or strong ion approach utilizes either original Peter Stewart equations or its simplified version deri-ved by Peter Constable (1,2). The strong ion and bicarbonate-centered approaches are qualitative-ly identical even in the presence of non-bicarbo-nate buffers. The traditional approach for inter-preting acid-base disorders developed from the pioneering work of Henderson and Hasselbalch is still the most widely used in clinical practice. One of the advantages of this approach is that it is relatively easy to understand and to apply in clinical situations (3). The strong ion model of-fers a novel insight into the pathophysiology of mixed acid-base disorders and is mechanistic (4). The newest approach has steadily gained accep-tance especially among critical-care physicians and anesthesiologists (5,6). In this review we will discuss the advantages and disadvantages of Stewarts method of acid-base balance compa-red with traditional bicarbonate-based approach, what will help better understanding in the com-plexity of acid-base balance.

    THE TRADITIONAL APPROACH

    In the early 1900s sufficient laboratory and observational evidence had been accumulated to define the influence of carbon dioxide on pH and to suggest a role for serum bicarbonate in charac-terizing acid-base disorders. Henderson recogni-zed that carbon dioxide and bicarbonate were key elements of carbonate mass action, as shown by his famous equilibrium equation (7). Hasselbalch reformulated this equation by introducing the ne-gative logarithmic pH notation and by applying Henrys law to generate the pCO2 term. A buffer is substance that has ability to bind or release hy-drogen ions (H+) in solution, thus keeping the pH of the solution relatively constant despite the ad-dition of considerable quantities of acid or bases (8). Bicarbonate is the most important buffer in biological system at constant pCO2, and Hender-son-Hasselbalch equation provides a simple rela-tionship among the respiratory parameter pCO2,

    the non-respiratory parameter bicarbonate, and overall acidity parameter, pH (9). But a change in bicarbonate concentration does not reflect the total amount of non-carbonic acid or base, beca-use there are non-bicarbonate buffers, especially albumin and hemoglobin (10). Even more impor-tant is the fact that bicarbonate concentration is not independent of variation in pCO2. As pCO2 increases carbonic acid is buffered by non-bicar-bonate buffers and the bicarbonate concentration increases. An elevated bicarbonate concentration may therefore erroneously be interpreted as a me-tabolic alkalosis when respiratory acidosis is the cause (11). So, this equation lists pCO2 and bi-carbonate as independent predictors of pH while, in fact, these variables are interdependent. Con-sequently, this equation merely serves as a des-cription of a patients acid-base state but does not provide insight into the mechanism of the patient s acid base disorder (12,13).The traditional approach to clinical acid-base in-terpretation (bicarbonate-centered approach) is based on Lowry-Bronsted theory, wherein acids are defined as substances capable of donating protons, and the centrality of the bicarbonate buffer system in whole body acid-base home-ostasis, given that it is composed of a volatile and nonvolatile buffer pair. Two bicarbonate-centered approaches evolved: the comparative HCO3/pCO2 approach (14) and the base exce-ss (BE) approach (15).

    The comparative HCO3/pCO2 approach

    Qualitatively, this approach starts with the assumption that the components of the HCO3-CO2 equilibrium reaction are in equilibrium with non-bicarbonate buffers (albumin, phosphate, he-moglobin) (3). Two bicarbonate-centered approaches evolved: the comparative HCO3/pCO2 approach (14) and the base excess (BE) approach (15).The HCO3/pCO2 approach has been critici-zed (2,11) for being qualitative in nature and incapable of quantifying acid or base loads that result in metabolic acid-base disorders. In par-ticular, because body compartments consist of multiple buffers, it is argued that HCO3 buffer is only one of several buffers that were protonated by an H+ load (16). Therefore, the HCO3 wo-uld underestimate the actual total body acid bur-

  • 9Todorovi et al. Assessment of acid-base analysis

    den as, for example, in a patient with keto-aci-dosis (17). Additional criticism is the fact that a component of the change in the HCO3 is due to a shift in the HCO3-CO2 equilibrium reaction as a result of the compensatory ventilatory res-ponse (altered pCO2) that occurs in patients with metabolic acid-base disturbances (18). Moreo-ver, the compensatory ventilatory induced al-teration in the pCO2 causes a change in renal NH4 and HCO3 and titrate-able acid excretion. This results in a further change of the [HCO3] (which is independent of the acid-base load and independent of the change due to the shift in the HCO3-CO2 equilibrium reaction) (19, 20). Fi-nally, a disadvantage is that the HCO3/pCO2 ratio expected in acute respiratory acid-base disorders depends on the number of proton bin-ding sites on non-bicarbonate buffers (albumin, hemoglobin, phosphate) (2). Thus, the bicarbo-nate concentration may be used as a screening parameter of a non-respiratory acid-base distur-bance when respiratory disturbances are taken into account (2).One possibility to solve this problem is to mea-sure the bicarbonate concentration at a standard condition (when pCO2 is 40 mm Hg) - value that we call standard bicarbonate, or to use the sum of bicarbonate and non-bicarbonate buffer ani-ons value that we call buffer base (changes in pCO2 would not affect the buffer base concen-tration as the rise in bicarbonate concentration associated with a rise in pCO2 is matched with a fall in concentration of other buffer anions) (11). Disagreement about the best parameter to descri-be acid-base balance in the body has dominated this area of physiology for more than three de-cades (19). The Henderson-Hasselbalch equation does not satisfactorily explain why the apparent value of pK1 in plasma depends on pH, protein concentration and Na+ concentration as well as the fact that only a non-linear relationship exists between log pCO2 and pH in vivo (markedly in acidic plasma) (2, 21). So, this equation (especi-ally HCO3/pCO2 ratio) is criticized because of its qualitative nature and impossibility to qu-antify the acid or base excess that exists in acid base balance disorders (20). So, this approach can only be accurately applied to human plasma at approximately normal pH, protein concentrati-on, and Na+ concentration (22).

    Base excess (BE) approach

    Dissatisfactions with the Henderson-Hasselbalch approach prompted Singer and Hastings to pro-pose in 1948 that plasma pH may be determined by two independent factors, pCO2 and net strong ion charge as the difference between all of the cations (termed total base) and anions (termed total fixed acid). They introduced the concept of buffer base (BB), (23) as the sum of all plasma buffer anions, i.e. bicarbonate plus the non-vo-latile weak acid buffers (albumin, phosphate) (3,24). These parameters (standard bicarbonate and buffer base) suggested as measures of me-tabolic acid-base disturbance are now obsolete.It is shown that a change in buffer base corres-ponds to a change in the metabolic component of acid-base balance, and yields the base excess (BE) methodology (25-28). Base excess is tra-ditionally calculated from Van Slyke equation as developed by Siggard-Anderson (29). In the late 1950s, Siggaard-Andersen and colleague, at a fixed temperature and partial pressure of carbon dioxide, measured the plasma bicarbonate con-centration and compared the difference between this value and a reference (30). When corrected by a constant, this difference yields the base exce-ss (BE) as a more sensitive measure of metabolic imbalance. Clinically, this base excess represents the amount of acid per unit volume that must be added to achieve a normal pH (27). Blood base excess was introduced to replace plasma [HCO3] with a measure of the metabolic component that is independent from the respiratory component, and incorporates the effect of hemoglobin as a buffer (30). Base excess represents the amount of acid or alkali that must be added to 1l of oxy-genated blood exposed in vitro to a pCO2 of 40 mm Hg to achieve the average normal pH of 7.40 (30). Acid is required when blood pH is higher than 7.40 (positive BE or base excess), where-as alkali is needed when blood pH is lower than 7.40 (negative BE or base deficit). Under normal conditions, the average blood BE is zero (11). Criticism of this method followed soon (18, 31, 32). For example, the laboratory BE value re-presents the net effect of all metabolic acid-ba-se abnormalities. Therefore, the effect of coexi-sting metabolic acidosis and alkalosis may lead to falsely suggest that no acid-base abnormality exists. Furthermore, this BE does not propose

  • Medicinski Glasnik, Volume 12, Number 1, February 2015

    10

    an etiology for the acid-base disorder once an abnormality is discovered (33). During in vitro blood titration, any CO2 induced increase in plasma concentration of bicarbonate is attended by an equivalent decrease in the anionic charge of non-bicarbonate buffers (mainly hemoglobin). This comes from the binding of H+ released from carbonic acid, and as a result blood base excess remains constant. Base excess becomes contro-versial because in vivo base excess is altered by purely respiratory changes. This occurs because blood freely exchanges ions with interstitial flu-id, which contains little or no protein buffer. The-refore, as PaCO2 changes in vivo, whole blood base excess changes measurably as bicarbonate and other ions equilibrate between blood and the interstitial space. Thus, primary PaCO2 change in living organism causes base excess to move in the opposite direction, despite demonstrable in vitro CO2 invariance. When the PaCO2 is varied in vivo as a result of hypoventilation or hyper-ventilation, blood base excess does not remain constant because a concentration gradient for bi-carbonate develops between blood and interstitial compartment. Accordingly, bicarbonate is remo-ved from the plasma into the interstitial fluid in hypercapnia resulting in a negative base excess, whereas bicarbonate is added to plasma from the interstitial fluid in hypocapnia causing positive base excess (1,5,34). When the pCO2 is varied in vivo by CO2 inhalation or hyperventilation, not only blood but all extracellular fluid is equilibra-ted with the new pCO2. When pCO2 increases pH tends to decrease more in the poorly buffered interstitial fluid than in the well buffered blood. H+ therefore tends to diffuse from the interstitial fluid into the blood where they are buffered in the erythrocytes. This addition of H+ to the blood is registered as a fall in whole blood base exce-ss, while the plasma base excess rises slightly. The actual ionic movements involve a diffusion of bicarbonate ions from the erythrocytes to the plasma and interstitial fluid in exchange for chlo-ride ions. However, the base excess of the total extracellular fluid remains constant, during acute pCO2 changes in vivo. It is not possible to obtain a sample of average extracellular fluid (including erythrocytes). However, a blood sample diluted threefold (1 + 2) with its own plasma may ser-ve as a model of extracellular fluid. This pitfall was addressed by introducing the extracellular

    base excess or standard base excess by Sigga-ard-Andersen (11), as a measure of the metabolic component that is modeled by diluting the blood sample threefold with its own plasma or estima-ted by using the blood base excess at a hemo-globin concentration of 50 g/l. Thus, standard base excess or extracellular base excess is mo-deled from the existing PaCO2 and pH, at a he-moglobin concentration of approximately 50 g/l, to replicate the mean extracellular hemoglobin concentration and does appear to have acceptable CO2 invariance in vivo, although it is less than perfect. The base excess equation was modified to standardize the effect of hemoglobin on CO2 titration in order to improve the accuracy of the base excess in vivo. Base excess of such model of the extracellular fluid may be calculated using the Van Slyke equation and now it represents the most relevant measure of a metabolic acid-base disturbance. Standard base excess is therefore ro-ughly the corrective dose of sodium bicarbonate in mmol per liter of extracellular fluid. Currently, many blood gas analyzers calculate standard base excess from measured pH, pCO2 and hemoglo-bin. Modern pH-blood gas analyzers calculate the extracellular base excess and present the re-sult with the same ease as they present the actual bicarbonate concentration. Determination requ-ires an arterial blood sample and a modern pH blood gas analyzer. Total CO2 (bicarbonate) mea-sured in venous plasma using an electrolyte anal-yzer or multi-purpose chemical analyzer may be used as screening parameter in patients without respiratory disorders. Base excess is numerically identical with the delta buffer base of Singer and Hastings: the change in buffer base from the va-lue at pH 7.4 and pCO2 40 mm Hg. However, standard base excess still yields results that are slightly unstable as pCO2 changes. Furthermore, the equation assumes normal concentration of non-bicarbonate buffers (albumin and phospha-te), and when albumin or phosphate is decreased, a common scenario in the critically ill patients, standard base excess will result in even more instability (4,11,23,35). Even that standard base excess became (and remains) an optional compu-tation that could be printed by most commercial blood gas analyzers, Bill Schvartz and Arnold Relman in Boston 1967 continued to advocate for the use of the actual bicarbonate concentra-tion in assessing a metabolic component of acid-

  • 11

    Todorovi et al. Assessment of acid-base analysis

    base status. Discussions in letters to the editors of several journals were called the great trans-Atlantic acid-base debate by John Bunker, and Boston and Copenhagen school were unreconci-led (4,11,23,35). So, in a traditional approach the metabolic component of acid-base physiology is based on the analysis of plasma concentrations of bicarbonate and standard base excess. Both are usually used in clinical practice and their calcula-tions are included in all blood gas analyzers. Despite these complexities, the comparative HCO3/pCO2 and BE approaches evolved hi-storically as two alternative bicarbonate-cente-red approaches for diagnosing clinical acid-base disturbances (2,36).

    The anion gap method

    An additional diagnostic contribution in asse-ssing metabolic component of acid-base anal-ysis, the anion gap method, was eventually introduced. The anion gap is defined as the di-fference between unmeasured plasma anions and the unmeasured plasma cations. A normal anion gap is 124 mmol/l. Since normally the total unmeasured anions exceed the total unme-asured cations, there is an anion gap. Under nor-mal conditions, the bulk of the serum anion gap (approximately 80%) is due to the sum of the ani-onic charges on circulatory proteins (albumin is the most abundant of circulating proteins). The charge on albumin at pH 7.4 contributes ~66% of the total net charge calculated by the anion gap, with the remainder composed of phospha-te, urate, lactate, ketone bodies, and sulfate (37). Usually proteins behave as anions, contributing about 13 mmol/L to the unmeasured anion pool (pH-independent protein charge is 3.7 mmol/L, pH-dependent protein charge is 10.3 mmol/ L, pH-dependent phosphate charge is 1.0 mmol/L, the net protein charge of human plasma = 3.7 + 10.3 1.0 = 13 mmol/L) (37). Therefore, changes in the concentration of serum albumin would be expected to alter the serum anion gap. For each 10 g/L decrement in the serum concentration of albumin, the serum anion gap was decreased by 2.5 mmol/L, and needed to be corrected to com-pensate for abnormal albumin concentrations (thus, for example, significant ketoacidosis, co-uld be missed in a diabetic patient with hypoalbu-minemia): corrected anion gap = observed anion

    gap 0.25 x ([normal albumin g/L] - [abnormal albumin g/L]), (38). This corrected anion gap can unmask an organic acidosis that was previously undetected in the setting of hypoalbuminemia (33). So, changes in the anionic effect of albumin will alter both the anion gap and the base excess. The presence of organic acid (XAH), which dissociates to form the anionic species XA whi-le consuming bicarbonate results in increased unmeasured anions (UA) and a subsequently in-creased anion gap. Therefore, an increase in the calculated anion gap compared to the institutio-nal reference, suggests the presence of an organic metabolic acidosis. Thus, when anion gap acido-sis exists, the increase in the anion gap should qualitatively mirror the fall in bicarbonate. Dis-ruption of this expected relationship is indicative of certain mixed acid-base disorders (39). The small size of the anion gap is most likely to re-sult from a reduced concentration of the normal unmeasured anion albumin in critically ill pa-tients. Globulins do not have a significant charge contribution compared to albumin since their pKa is much greater than plasma pH (Ka is the effec-tive equilibrium dissociation constant for plasma weak acids). As myeloma proteins have isoelec-tric points >7.4 they become positively charged in the serum and behave as cations. In this way they may lead to a reduced anion gap by creating an excess of positively charges ions. That has to be counterbalanced by an increase in anions, ma-inly chloride. This explains why in about 30% of myeloma or gammapathies the anion gap could be

  • Medicinski Glasnik, Volume 12, Number 1, February 2015

    12

    cal pH range (Na+, K+, Ca2+), nonvolatile proton donor/acceptors which transfer H+ within the physiological pH range (albumin, phosphate, he-moglobin, metabolizable organic compounds), and the volatile bicarbonate-CO2 buffer system composed of CO2, HCO3, H2CO3, and CO3

    2-. According to requirements of electro neutrality, the law of conservation of mass and certain equ-ilibrium constants, Stewart solved a fourth-order polynomial equation for calculating the H+ con-centration. His analysis did not depart from the traditional approach to acid-base chemistry as a result of the equations that are derived. Pivotal to the Stewart formulation was the categorization of certain species as being dependent or indepen-dent variables in relationship to their purported role in determining and modifying the H+ con-centration of solution. At first, H+, OH-, HCO3, and CO3

    2- were categorized as dependent varia-bles (the mass balance of these species in a so-lution or specific body fluid compartment could not per se affect the H+ concentration). Finally, he contended that H+ concentration was a functi-on of three variables: strong ion difference (SID) (the difference in the net charge of fixed cations and anions fully dissociated in solution), parti-ally dissociated weak acids (albumin, phosphate) (ATOT), and the partial pressure of carbon dioxi-de (pCO2) of the solution (2). According to the principle of electro neutrality, SID is balanced of the weak acids (albumin, phosphate) and CO2. Therefore, SID can be defined either in terms of strong ions or in terms of the weak acids and CO2 offsetting it. Of note, the SID defined in terms of weak acids and CO2, termed as the effective SID and is identical to the buffer base term coined by Singer and Hastings over half a century ago. So, changes in standard base excess also represent changes in SID (42). Stewart hypothesized that water dissociates into H+ and OH- to a greater or a lesser extent when (SID), (ATOT ) or pCO2 change. Aqueous soluti-ons contain a virtually inexhaustible source of H+. Although pure water dissociates only slightly into H+ or OH-, electrolytes and CO2 produce powerful electrochemical forces that influence water dissociation. SID has a powerful electro-chemical effect on water dissociation, and hence on H+ concentration. As SID becomes more po-sitive, H+, a weak cation, decreases - (and pH

    increases) in order to maintain electrical neutra-lity. Strong ions cannot be created or destroyed to satisfy electro neutrality but H+ ions are ge-nerated or consumed by changes in water dis-sociation (30). This method emphasizes the role of water dissociation as a proton source and the association of water to consume protons as the driving force behind changes in blood pH. The relative balance of plasma positive and negati-ve charges, including those on serum proteins, most notably albumin, drives water dissociation and association by the law of mass action (43). Stewart approach puts water dissociation at the center of the acid-base states of body fluids. pH of a body fluid is a function of water dissociation modified by pCO2, other weak acids and certain electrolytes (5,44). To date there have been no empirical observations that confirm water disso-ciation as the mechanism whereby (SID), (ATOT) or pCO2 affect pH. Thus, the disadvantage of Si-ggaard-Andersen approach is that it implies ad-ding or introducing H+ to the solution, which is impossible. The more general Stewarts approach explains the acid base variations over more valu-able physical basis. This physicochemical appro-ach can be helpful in understanding mechanisms barely understandable using traditional approa-ch (45). The Stewart approach is a very general physicochemical method that uses charge and mass balance to deduce an expression for proton concentration. Similarly, the base excess method is another very general physicochemical approa-ch, but one that uses proton balance to calculate changes in proton concentration by using the Van Slyke equation (46). Thus, Stewart challenged the traditional bicar-bonate-based method of diagnosis and treating acid-base disorders and proposed an approach based primarily on charge differences between strong cations and anions. He suggested that three independent factors, the pCO2, the strong ion difference, and the total non-volatile weak acid concentration may be used to analyze the causes of acid-base disorders. These three fac-tors control other variables including hydrogen ion concentration and bicarbonate concentration which are dependent variables. Increases in pCO2 and total weak acid concentration increase aci-dity. Decreases in strong ion difference increase acidity. This approach may better explain the me-

  • 13

    Todorovi et al. Assessment of acid-base analysis

    chanism of acid-base physiology and disorders than the Henderson-Hasselbalch approach (44). On the basis of Stewart s definition H+ and bicar-bonate are dependent variables whose concentra-tions are determined by three independent varia-bles, namely SID, pCO2 and ATOT. SID therefore can be calculated as the difference between fully dissociated cations and anions which do not par-ticipate in proton transfer reactions (aprote ions - neither donate or accept H+) and by the principle of electro neutrality this difference is equal to the sum of bicarbonate and the non-bicarbonate ani-ons, which represent total charges contributed of all non-bicarbonate buffers, primarily, albumin and phosphate, and in whole blood, hemoglobin. SID is, therefore, the same as buffer base concept introduced by Singer and Hastings more than 5 decades ago.

    Apparent SID and effective SID

    When abnormal anion is present, a gap will appear between SID calculated by the differen-ce between strong ions (the so-called apparent SID) and calculated by the addition of bicarbo-nate and non-bicarbonate buffers (so called ef-fective SID). This difference, named strong ani-on gap (SIG), is a marker for the presence of an abnormal anion (23). The SIG indicates the pre-sence of unmeasured strong anions if its value is positive, the normal value of the SIG is zero. The SIG is similar in concept to the anion gap, once the latter has been corrected for anionic contri-bution of albumin and phosphate (3). A positive SIG value represents unmeasured anions (such as ketoacids, urate, sulfate, citrate, pyruvate, acetate and gluconate) that are present in the blood.Apparent SID is referred to as apparent (SIDa-pp) with the understanding that some unmeasured ions might be also present. It represents the diffe-rence between the charge of measured strong ca-tions (sodium, potassium, calcium, magnesium) and strong anions (chloride, lactate, sulfate, keto-acids, nonesterified fatty acids and many others) that are completely dissociated in biological so-lutions. Currently, (SID) is routinely measured from [Na+] + [K+] + [Ca2+] + [Mg2+] - [Cl-] - [lac-tate-]. It is always positive in plasma and in heal-thy humans - its value is 40 to 42 mmol/L, while it is often quite different in critically ill patients. The effective SID (SIDeff) represents the effect of

    the corrected pCO2 and the weak acids, albumin and inorganic phosphate, on the balance of electri-cal charge in plasma, where HCO3 is in mmol/L, albumin in g/l and phosphate in mmol/L. All the independent variables are present in millimolar concentrations and their interaction with water dictates the amount of free H+, the concentration of which is in order of nanomoles. The difference between the calculated apparent SID and effective SID constitutes the strong ion gap, SIG. In healthy humans, the SIG should be theoretically equal to zero (electrical charge neutrality). If this is not the case, there must be unmeasured charges to explain this ion gap. A positive SIG value represents unmeasured anions (such as ketoacids, urate, sulfate, citrate, pyruva-te, acetate and gluconate) that are present in the blood and account for the measured pH, the me-asured levels of strong and weak ions, and the need to maintain electro neutrality (46). Unlike the classic parameters used to calculate the ani-on gap, in calculating SIG, the effect of albumin, phosphate and lactate is subtracted out. There-fore, SIG compared with the anion gap is caused by shorter list of unmeasured ions including ke-tone bodies, sulfate and uraemic anions (2).Calculating the effective SID takes into account the role of weak acids (carbon dioxide, albumin and phosphate) in the balance of electrical char-ges of plasma water (47). Once weak acids are qualitatively taken into account, the difference between apparent and effective SID should be zero, unless there are unmeasured charges (ani-ons). Such charges are then described by the strong ion gap, SIG (SIG = apparent SID ef-fective SID). Bicarbonate is considered separa-tely because this buffer system is an open system in arterial plasma. Rapid changes in pCO2 and hence arterial bicarbonate concentration can be rapidly induced through alteration in respiratory activity. In contrast, the non-bicarbonate buffer system is a closed system containing relative-ly fixed quantity of buffers. Nonvolatile buffer ion (A-) represents a diverse and heterogeneous group of plasma buffers consisting primarily of dissociable imidazole and -amino groups on plasma proteins with a smaller contribution from phosphate-containing weak acids and citrate.On the basis of information stated above, plasma contains three types of charged entities: SID+,

  • Medicinski Glasnik, Volume 12, Number 1, February 2015

    14

    HCO3 and A-. The requirement for electro neu-

    trality dictates that at all times [SID+] equals the sum of [HCO3] and nonvolatile [A

    -], such that [SID+] - [HCO3] - [A

    -] = 0. (The dissociation re-action for a weak acid-conjugate base-pair, HA and A-, is HA =H++A-) (48).

    Partially dissociated weak acids (albumin, phos-phate), (ATOT)

    Plasma proteins provide the major contribution to ATOT, and therefore, plasma protein concen-tration independently affects acid-base balance. One major departure from the traditional appro-ach is classification of acid-base disorders as a result of alteration in ATOT. ATOT, representing all non-bicarbonate buffers pairs (HA + A-), is made up of charges contributed primarily by serum proteins (mainly albumin) with phosphate and other buffers playing a minor role. The sum of [HA] and [A-] (called ATOT by Stewart) therefore remains constant through conservation of mass. The normal value of the total negative charges on plasma non-bicarbonate buffers is [ATOT] = [A

    -albu-

    min] + [A-globulin] + [A

    -phosphate] = 16.6 mmol/L (49).

    On the basis of this classification, an increase in serum protein would result in metabolic acidosis while a decrease would cause metabolic alkalo-sis. Siggaard-Andersen and Fogh-Andersen have suggested that changes in protein concentration should not be considered as acid-base disorders, and have considered the Stewart approach pro-blematic in that regard (11). Dissatisfaction with the Henderson-Hasselbalch approach prompted Singer and Hastings to propose in 1948, that pla-sma pH was determined by two independent fac-tors, pCO2 and net strong ion charge, equivalent to the strong ion difference (SID+). Stewart later proposed that third variable, the total plasma con-centration of nonvolatile weak acids (ATOT), also, exerted an independent effect on plasma pH. Thus, by combining equations for conservation of charge, conservation of mass, and dissociation equilibrium reactions, Stewart developed a pol-ynomial equation, relating H+ concentration [H+] to three independent variables (pCO2, [SID

    +] and [ATOT]) and five constants (48). These three in-dependent variables may change the hydrogen concentration in water (i.e. the acid-base equili-brium). The strong ion difference is regulated by the kidney, weak acid concentration primarily by

    liver, and pCO2 by lung (48). An independent va-riable is defined as one that influences the system but is not influenced by the system. The term sy-stem refers to any single aqueous compartment (i.e. plasma).The Stewarts approach states that pH is prima-rily determined by several independent varia-bles (which change primarily and independently of one another: by pCO2, strong ion difference (SID) and nonvolatile weak acids). This physi-cochemical approach might identify altered indi-vidual component of complex acid-base abnor-malities and provide insights to their underlying mechanisms (35).

    The relation between strong ion difference (SID) and standard base excess (SBE)

    The SIDa must be counterbalanced by an equal and opposing charge termed the effective strong ion difference (SIDe) (normal approximately -40 to -42 mmol/L). The SIDe negative charge prin-cipally stems from the dissociated moieties of plasma proteins (-78% albumin) and phosphate (-20%). The sum of these weak acids is known as ATOT since they exist in a dissociated form (A

    -), as well as an associated form (AH). When the SIDa and SIDe are equal the plasma pH is exactly 7.4 at a pCO2 of 40 mm Hg. Bicarbonate is depen-dent variable and does not determine the pH (47). According to the principle of electro neutrality, SIDa is balanced of the weak acids and CO2, such that SID can be defined either in terms of strong ions (SIDa) or in terms of the weak acids and CO2 (SIDe) offsetting it. Of note, the SID defined in terms of weak acids and CO2, which has been su-bsequently termed the effective SID is identical to the buffer base term coined of Singer and Ha-stings over half a century ago. So, changes in the standard base excess also represent changes in SID. The difference in electrical charge between strong cations and strong anions is called strong ion difference. In normal plasma this amount is about 42 mmol/L. Indeed, to reach the electro neutrality, 42 mmol/l of negative charged ions, are required. These are basically the bicarbonate (HCO3) and negative charged form of weak acids (A-), primarily albumin plus an extremely small amount of hydroxyl (OH-). The sum of [HCO3] + [A-] which equals the strong ion difference was called buffer base by Singer and Hastings, and

  • 15

    Todorovi et al. Assessment of acid-base analysis

    later by Siggaard-Andersen. Indeed, a big dif-ference between the traditional approach (SBE) and the Stewarts approach is that the first con-siders what happens inside buffer base domain. As an example, in the traditional model (SBE) the normal buffer base (42 mmol/l, as the normal [SID]) may decrease for 10 mmol/L if the (A-) and (HCO3) are consumed by adding 10 mmol/L of H+. In this case the actual buffer base is 32 mmol/L and the difference between the actual bu-ffer base and the ideal buffer base is equal to -10 mmol/L, and is called base excess. In the Stewart model the same problem is con-sidered from another point of view. If a strong ion is added to the system the ratio between strong cations and strong anions will change. As an example, by adding 10 mmol/l of lactate the strong ion difference decreases from 42 mmol/L to 32 mmol/L. The space available for A- and HCO3 and OH

    - decreases, indeed part of A- will become AH, part of HCO3 will become H2CO3 and part of OH- will become H2O. As the product of H+ and OH- is constant, a decrease of OH- will correspond to an increase of H+. i.e. acidosis (45). So, the disadvantage of standard base excess approach is that it implies adding or introducing H+ to the solution, which is impossible. The more general Stewarts approach explains the acid-ba-se variations on a more valuable physical basis. So, the physicochemical Stewarts approach can be helpful in understanding mechanisms barely understandable using traditional approach.The mystery of dilutional acidosis during re-pletion of extracellular fluid deficit can be better explained by the modern approach. The mechani-sm is obviously not bicarbonate dilution as expla-ined by the traditional approach (otherwise why would the proton donors not be diluted at the same time?). Therefore, from Stewarts perspective an alternative explanation for dilutional acidosis nee-ded to be developed. His mechanistic explanation is based on strong ions and the maintenance of electro neutrality. It was believed that positive or negative charges (i.e. changes of the concentrati-ons of strong ions) influence the dissociation of water (50). In the context of dilutional acidosis, this means dilution of plasma (which has a positi-ve SID of 39 mmol/l) by water or another solution with SID of zero decreases the SID, i.e. diminishes the surplus of positive charges. However, the de-

    crease in SID demands compensation by a positive charge. This is suggested by increased water dis-sociation with generation of a positively charged H+. This newly generated H+ then causes acidifi-cation of the solution, i.e. dilutional acidosis (50). Interestingly, hypertonicity makes solutions more acidifying, as more water is drained from the intra-cellular space, which ultimately contributes to the final equilibrium (51).

    Changes in SID, SIG and ATOT in acid base disorders

    Respiratory disorders, in the modern approach as in the traditional approach, are due to change in pCO2, whereas metabolic disorders are due to al-terations in either SID or ATOT. SID is decreased in metabolic acidosis and increased in metabolic alkalosis. By calculating SIG, one can further cla-ssify metabolic acidosis. In hyperchloremic me-tabolic acidosis both effective and apparent SID decreases equally, as the increase in chloride is counterbalanced by an equal decrease in the bicar-bonate concentration. SIG therefore remains at or near zero. In anion gap metabolic acidosis, appa-rent SID does not change (as chloride concentrati-on is unchanged), but effective SID decreases (as a result of a decrease in bicarbonate concentrati-on) and SIG therefore becomes positive, reflecting high levels of unmeasured anions such as lactate and ketoanions (52). One major departure from the traditional approach is classification of acid-base disorders as a result of alteration in ATOT. On the basis of this classification, an increase in serum protein would result in metabolic acidosis and a decrease, metabolic alkalosis (23).

    CONCLUSION

    Three distinct approaches are currently used in assessing acid-base disorders, the traditional or physiological approach pioneered by Van Slyke, the base-excess method, developed by Astrup, and the physicochemical approach, proposed by Stewart. The last and newest approach has stea-dily gained acceptance especially among critical-care physicians and anesthesiologists (5).The traditional approach to interpreting acid-ba-se disorders developed from the pioneering work of Henderson and Hasselbalch and is still most wi-dely used in clinical practice. An advantage of this approach is that it is relatively easy to understand and to apply in common clinical situations (3).

  • Medicinski Glasnik, Volume 12, Number 1, February 2015

    16

    We conclude that the physiological or traditional approach remains the simplest, most rigorous, and most serviceable approach to assessing acid-base disorders. Clinically, the traditional approach is intuitive in nature and is supported by a large body of robust empirical observations. The true relevant acid-base quantities are the arteri-al pH, the arterial pCO2, and the extracellular base excess. Determination requires an arterial blood sample and a modern pH blood gas analyzer. To-tal CO2 (bicarbonate) measured in venous plasma using an electrolyte analyzer or multipurpose che-mical analyzer may be used as screening parameter in patients without respiratory disorders (101).The strong ion model offers a novel insight into the pathophysiology of a mixed acid-base dis-orders and is mechanistic (4). The traditional approach should be abandoned only if propo-nents of Stewarts approach could provide clear empirical observations supporting its superiority as a clinical tool in diagnosis and treating patients with acid-base disorders. The dependence of pK1 (carbonic acid) on pH and protein concentration is a major anomaly for the Henderson-Hasselbalch equation because the dissociation constant for equilibrium reactions should not be influenced by changes in reactants (hydrogen ion activity = pH) or by anything else (including protein) except temperature and ionic strength, the latter being determined primarily by the sodium concentration. The change in [SID+] from normal is equivalent to the base excess va-lue assuming a normal, nonvolatile buffer ion concentration (normal albumin, globulin and phosphate concentrations) (1).If serum total protein, albumin and phosphate concentrations are approximately normal, then

    acid-base status should be evaluated using blood pH, pCO2 and extracellular base excess concen-tration, which is the traditional Henderson-Ha-sselbalch approach. The presence of unidentified anions should be investigated by calculating ani-on gap. If albumin concentration is abnormal, the anion gap can be corrected based on the albumin concentration, as this corrected anion gap can unmask an organic acidosis that was previously undetected in the setting of hypoalbuminemia. However, if serum total protein, albumin and phosphate concentrations are markedly abnor-mal, as in critically ill patients, then acid-base status should be evaluated using blood pH, pCO2, measured [SID+] and ATOT. The presence of uni-dentified strong ions should be investigated by calculating the SIG. Traditional approach works well clinically and is recommended for use whe-never serum total protein, albumin and phospha-te concentrations are normal. Although Stewarts approach has been largely ignored by physiolo-gists, it is increasingly used by anesthesiologists and intensive care specialists, and is recommen-ded for use whenever serum total protein, albu-min or phosphate concentrations are markedly abnormal, as in critically ill patients.Although different in their concept, the traditio-nal and modern approaches can be seen comple-mentary giving, in principle, the same informati-on about the acid-base status.

    FUNDING

    This work was supported by grants 175081 from the Ministry of Education, Science and Techno-logical Development of the Republic of Serbia.

    TRANSPARENCY DECLARATION

    Competing interests: None to declare.

    REFERENCES

    1. Constable PD. Clinical assessment of acid-base sta-tus: comparison of the Henderson-Hasselbalch and strong ion approaches. Vet Clin Pathol 2000; 29:115-28.

    2. Kurtz I, Kraut J, Ornekian V, Nguyen MK. Acid-base analysis: a critique of the Stewart and bicar-bonate-centered approaches. Am J Physiol 2008; 294:F1009-31.

    3. Sirker AA, Rhodes A, Grounds RM, Bennett ED. Acid-base physiology: the traditional and the mo-dern approaches. Anaesthesia 2002; 57:348-56.

    4. Kellum JA. Disorders of acid-base balance. Crit Care Med 2007; 35:2630-6.

    5. Morgan TJ. The Stewart approach-one clinicians perspective. Clin Biochem Rev 2009; 30:41-54.

    6. Kishen R, Honore PM, Jacobs R, Joannes-Boyau O, De Waale E, De Regt J, Van Gorp V, Boer W, Spapen HD. Facing acid-base disorders in the third millenni-um the Stewart approach revisited. Int J Nephrol Renovasc Dis 2014; 7:209-17.

    7. Henderson LJ. Concerning the relationship between the strength of acids and their capacity to perserve neutrality. Am J Physiol 1908; 21:1739.

  • 17

    Todorovi et al. Assessment of acid-base analysis

    8. Henderson LJ. The theory of elecroneutrality regu-lation in the animal organism. Am J Physiol 1907; 18:427-8

    9. Van Slyke DD. Studies of acidosis I. The bicarbona-te concentration of the blood plasma; its significance and its determination as a measure of acidosis. J Biol Chem 1917; 30:289-46.

    10. Hughes R, Brain MJ. A simplified bedside approach to acid-base: fluid physiology utilizing classical and physicochemical approaches. Anesthesia & Intensi-ve Care Medicine 2013; 14:445-52.

    11. Siggaard-Andersen O, Fogh-Andersen N. Base excess of buffer base (strong ion diference) as me-sure of a non-respiratory acid-base disturbance. Acta Anaesthesiol Scand 1995; 39(Suppl. 106):123-8.

    12. Hickish T, Farmery AD. Acid-base physiology: new concepts. Anesthesia & Intensive Care Medicine 2012; 13:567-72.

    13. Wilcox C, Cemerikic D, Giebisch G. Differential effects of acute mineralo- and glucocorticosteroid administration on renal acid elimination. Kidney Int 1982; 21:546-66.

    14. Schwartz WB, Relman AS. A critique of the para-meters used in the evaluation of acid-base disorders whole-blood buffer base and standard bicarbona-te compared with blood pH and plasma bicarbonate concentration. New Enlg Med 1963; 268: 1382-8.

    15. Siggaard-Andersen O. The acid-base status of the blood (4th Ed.). Baltimore, MD: William and Wil-kins, 1974.

    16. Hubble SMA. Acid-base and blood gas analysis. Anesthesia & Intensive Care Medicine 2007; 8:471-3.

    17. Severinghaus JW. Case for standard-base excess as the measure of nonrespiratory acid-base imbalance. J Clin Monit 1997; 7:276-7.

    18. Siggaard-Andersen O, Wimberley PD, Fogh-Ander-sen N, Gothgen IH. Measured and derived quantities with modern pH and blood gas equivalent circulati-on algorithm with S4 equations. Scand J Clin Lab Invest 1988; 48:7-15.

    19. Kurtz I. Acid Base Case Studies. 2006. Victoria, Ca-nada: Trafford.

    20. Madias MA, Schwartz WB, Cohen JJ. The maladap-tive renal response to secondary hypocapnia during chronic HCl acidosis in the dog. J Clin Invest 1997; 60:1393-401.

    21. Cemerikic D, Wilcox C, Giebisch G. Intracellular potential and K+ activity in rat kidney proximal tubu-lar cells in acidosis and K+ depletion. J Membr Biol 1982; 69:159-65.

    22. Berend K. Acid-base pathophysiology after 130 ye-ars: confusing, irrational and controversial. J Nep-hrol 2013; 26:254-65.

    23. Rastegar A. Clinical utility of Stewarts method in diagnosis ans management of acid-base disorders. Clin J Am Soc Nephrol 2009; 4:1267-74.

    24. Wooten EW. Analytic calculation of physiological acid-base parameters in plasma. J Appl Physiol 1999; 86:326-34.

    25. Astrup P, Jorgensen K, Siggaard-Andersen O. Acid-base metabolism: New approach. Lancet 1960; 1: 1035-9.

    26. Siggaard-Andersen O. The pH-log PCO2 blood acid-base nomogram revised. Scand J Clin Lab Invest 1962; 14:598-604.

    27. Grogno AW, Bzles PH, Hawke W. An in vivo repre-sentation of acid-base balance. Lancet 1976; 1:499-500.

    28. Severinghaus JW. Acid-base balance nomogram a Boston-Copenhagen detente. Anesthesiology 1976; 45:539-41.

    29. Siggaard-Anderson O. The Van Slyke equation. Scand J Clin Lab Invest 1977; 146:15-20.

    30. Kellum JA. Determinants of blood pH in health and disease. Crit Care 2000; 4:6-14.

    31. Gonzales NC, Clancy RL. Intracellular pH regulati-on during prolonged hypoxia in rats. Respir Physiol 1986; 65:331-9.

    32. Roos A, Thomas LJ. The in-vitro and in-vivo carbon dioxide dissociation curves of true plasma. Anesthe-siology 1977; 28: 1048-63.

    33. Fidkowski C, Helstrom J. Diagnosis metabolic aci-dosis in the critically ill: bridging the anion gap, Stewart and base excess methods. Can J Anest 2009; 56:247-56.

    34. Greenbaum J, Nirmalan M. Acid-base balance: Stewarts physicochemical approach. Curr Anae-sthesia Crit Care 2005; 16:133-5.

    35. Dubin A, Menises MM, Masevicius FD, Mosein-co MC, Kutscherauer DO, Ventrice E, Laffaire E, Estenscoro E. Comparison of three different met-hods of evaluation of metabolic acid-base disorders. Crit Care Med 2007; 35:1264-70.

    36. Geoffrey C, Aiken A. History of medical understan-ding and misunderstanding of acid-base balance. J Clinical Diagnostic Research 2013; 7:2038-41.

    37. Staempfli H, Constable PD. Experimental determi-nation of net protein charge and A tot and K a of nonvolatile buffers in human plasma. J Appl Physiol 2003; 95:620-30.

    38. Kraut JA, Madias NE. Serum anion gap: its uses and limitations in clinical medicine. J Am Soc Nephrol 2007; 2:162-74.

    39. Emmett M. Anion gap interpretation: the old and the new. Nat Clin Pract Nephrol 2006; 2:4-5.

    40. Story DA. Bench-to-bedside review: a brief histor of clinical acid-base. Crit Care 2004; 8:253-8.

    41. Correy HE. Stewart and beyond: new models of acid-base balance. Kidney Int 2003; 64:777-87.

    42. Kellum JA. Clinical review: Reunification of acid-base physiology. Crit Care 2005; 9:500-7.

    43. Kaplan LJ, Kellum JA. Initial pH, base deficit, lacta-te, anion gap, strong ion. Crit Care 2004; 32:1120-4.

    44. Smuszkiewicz P, Szrama J. Theoretical principles of fluid management according to the physicochemi-cal Stewart approach. Anaesthesiol Intensive Ther 2013; 45:99-105.

    45. Gattinoni L. Carlesso E, Maiocchi G, Polli F, Car-dingher P. Dilutional acidosis: where do the protons come from? Intensive Care Med 2009; 35:2033-43.

    46. Nguyen BV, Vincent JL, Hamm JB, Abalin JH, Carre JL, Nowak E, Ahmed MO, Arvieux CC, Gu-eret G. The reproducibility of Stewart parameters for acid-base diagnosis using two central laboratory analyzers. Anesth Analg 2009; 109:1517-23.

    47. Naka T, Bellomo R. Bench-to-bedside review: tre-ating acis-base abnormalities in the intensive care unit-the role of renal replacement therapy. Crit Care 2004; 8:108-14.

    48. Constable PD. A simplified strong ion model for acid-base equilibria: application to horse plasma. J Appl Physiol 1997; 83:297-311.

  • Medicinski Glasnik, Volume 12, Number 1, February 2015

    18

    49. Matousek S, Handy J, Rees SE. Acid-base chemistry of plasma consolidation of the traditional and mo-dern approaches from a mathematical and clinical perspectives. J Clin Monitor Comp 2011; 25:57-70.

    50. Doberer D, Funk GC, Kirchner K, Schneeweiss B. A critique of Stewarts approach: the chemical me-chanism of dilutional acidosis. Intensive Care Med 2009; 35:2173-80.

    51. Story DA, Morimatsu H, Bellomo R. Strong ions, weak acids and base excess: a simplified Fencl-Stewart approach to clinical acid-base disorders. Brit J Anaesthesia 2004; 92:54-60.

    52. Androgue HJ, Gennari FJ, Galla JH, Madias NE. Assessing acid-base disorders. Kidney Int 2009; 76:1239-47.

    Acido-bazne analize: poreenje tradicionalnog i modernog pristupaJasna Todorovi1, Jelena Neovi-Ostoji1, Aleksandar Milovanovi2, Predrag Brki3, Mihailo Ille4, Duan emeriki11Institut za patoloku fiziologiju, 2Institut za medicinu rada, 3Institut za fiziologiju, 4Hirurgija; Medicinski fakultet Univerziteta u Beogradu, Beograd, Srbija

    SAETAK

    Danas se koriste tri posebna pristupa u proceni acido-baznih poremeaja: tradicionalni ili fizioloki ili bikarbonatni pristup, metoda baznog ekcesa i moderni fizikalno-hemijski pristup predloen od Stewarta koji koristi razliku glavnih jona (naroito natrija i hlorida), koncentraciju neisparljivih slabih kiselina (osobito albumina) i parcijalni pritisak ugljen dioksida (pCO2) kao nezavisne varijable u analizi acido-baznog statusa. Tradicionalni pristup kojeg su predloili Henderson i Hasselbalch i metoda ba-znog ekcesa, uprkos brojnim nedostacima, jo uvek su najire korieni u klinikoj praksi. Ovaj pristup se naroito preporuuje u klinikoj praksi kada su koncentracije ukupnih proteina, albumina i fosfata u serumu normalne. Iako je Stewartova metoda bila uglavnom ignorisana od strane fiziologa, nju sve vie primenjuju anesteziolozi i lekari intenzivne nege prilikom leenja teko bolesnih pacijenata kod kojih su koncentracije ukupnih proteina, albumina i fosfata u serumu poremeene. Iako razliite u svom koncep-tu, tradicionalna i moderna metoda mogu se posmatrati kao komplementarni pristupi koji upotpunjuju informisanost o acido-baznom statusu.

    Kljune rei: acido-bazna ravnotea, anjonski zjap, razlika glavnih jona, bikarbonati, bazni eksces, neisparljive slabe kiseline, zjap glavnih jona