You are on page 1of 6

pubs.acs.

org/jchemeduc Article

Rising Atmospheric Carbon Dioxide Could Doom Ocean Corals and


Shellfish: Simple Thermodynamic Calculations Show Why
Todd P. Silverstein*
Cite This: J. Chem. Educ. 2022, 99, 2020−2025 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The inexorable rise in atmospheric carbon dioxide


impacts not only global warming but also the acidity of the ocean.
Increasing ocean acidity causes a decline in carbonate (as it is
Downloaded via UNIV DE ALICANTE on October 14, 2022 at 16:06:06 (UTC).

protonated), which in turn will negatively impact the ability of


calcifying marine organisms to build their calcium carbonate shells. A
simple set of equilibrium thermodynamic calculations appropriate for
first-year chemistry students shows that, based on carbonate
concentration predictions from reliable models, calcium carbonate
shell deposition could become nonspontaneous in the Southern Ocean
before the end of this century. The resulting decline in calcifying
plankton could in turn lead to ecosystem collapse in this region. For
physical chemistry students, effects of seawater salinity on equilibrium concentrations of carbonate can be explained using activity
coefficients and ionic strength. Also, differences in the carbon dioxide/carbonate system in polar vs tropical oceans are explained by a
combination of thermodynamics and the fact that carbon dioxide and carbonate concentrations are maintained out of equilibrium by
biological and physical processes.
KEYWORDS: High School/Introductory Chemistry, First-Year Undergraduate/General, Upper Division Undergraduate,
Environmental Chemistry, Physical Chemistry, Acids/Bases, Equilibrium, Thermodynamics, Applications of Chemistry,
Aqueous Solution Chemistry

■ INTRODUCTION
Many authors over the past few decades have stressed the
ecosystem, their health and abundance are of the utmost
importance.
importance of using real-world examples to connect chemistry The beauty of the marine carbon dioxide/calcium carbonate
to students’ everyday lives. Papers have been published system from a pedagogical perspective is that it is fairly simple,
describing “real-world” laboratory experiments and sympo- and the thermodynamic equilibrium calculations are appropriate
sia,1−5 case studies,6−9 semester-long courses featuring “real- for first-year chemistry students. As a result of a straightforward
world” chemistry,10−15 and even a “biologically relevant” four- thermodynamic analysis, students can understand not only the
year chemistry curriculum.16 differences between polar and tropical oceans, but also why
One environmental problem that has captured worldwide calcifying organisms could disappear entirely from the polar
attention is the global rise in atmospheric carbon dioxide due to oceans by the end of this century.
We describe here a case study that builds on three earlier
fossil fuel combustion. Although the impact of this and other
publications in this Journal5,24,25 exploring the connections
greenhouse gases on global warming is widely appreciated, the
between the greenhouse gas carbon dioxide and marine
influence of atmospheric carbon dioxide on marine chemistry
chemistry, concentrating on well-known thermodynamic
has not received quite as much attention. The ocean absorbs
relationships such as Hess’s law, standard Gibbs free energy vs
≈1/3 of the CO2 released into the atmosphere by human
equilibrium constant (ΔG° = −RT ln Keq), dependence of Gibbs
activity, which amounts to more than 6 million metric tons of
free energy on concentration (ΔG = ΔG° + RT ln Q), activity vs
“extra” CO2 absorbed per day!17,18 A number of reports have
concentration, and the Debye−Huckel limiting law. We distill all
described the dramatic negative impact of increased dissolved
CO2 on marine pH and carbonate concentrations. For example,
in a recent review, Zeebe stated “the ocean acidification event Received: February 17, 2022
that humans are expected to cause is unprecedented in the Revised: March 18, 2022
geologic past.”19,20 Lower carbonate concentration inhibits the Published: April 6, 2022
ability of marine animals to build calcareous (CaCO3)
shells.21−23 Because calcifying organisms (e.g., corals, shellfish,
foraminifera) are such an important part of the marine
© 2022 American Chemical Society and
Division of Chemical Education, Inc. https://doi.org/10.1021/acs.jchemed.2c00149
2020 J. Chem. Educ. 2022, 99, 2020−2025
Journal of Chemical Education pubs.acs.org/jchemeduc Article

of this into two worksheets (Supporting Information), one for second equilibrium in eq 4 is simply the sum ΔGa1 − ΔGa2 +
general chemistry students and one for physical chemistry ΔGsp. All three of these values under standard conditions
students. (concentrations 1 M) are readily available in introductory

■ ANALYSIS AND DISCUSSION


Carbon Dioxide, Sea Shells (CaCO3), and pH
chemistry textbooks: pKa1(H2CO3) = 6.365 (e.g., refs 26 and
27), pKa2(H2CO3) = 10.32 (e.g., refs 26 and 27), and
Ksp(CaCO3(s)) = 4.5 × 10−9 (ref 27). From these equilibrium
constants, we can calculate ΔG°net at 25 °C = 36.3(3) − 58.(90)
The carbon dioxide/calcium carbonate system essentially + 47.(64) = +25.(1) kJ/mol. Thus, under standard conditions,
comprises three reactions: calcium carbonate will not dissolve in the presence of 1 M
After CO2 dissolves in water to give CO2(aq), nucleophilic H2CO3 (CO2(aq)); furthermore, calcium carbonate precip-
attack by water gives carbonic acid; because these two are itation (i.e., shell-building) is spontaneous.
difficult to distinguish,24,25 H2CO3(aq) and CO2(aq) are
sometimes used interchangeably. Carbonic acid then ionizes Equilibrium Thermodynamic Calculations Under Ocean
(reaction 1, Figure 1) to bicarbonate and H+, raising the Conditions
If the calculation above was all there was to it, we would have no
need for the voluminous literature that has grown up to account
for marine chemistry. There are several important differences
between seawater and the standard solutions represented in the
pKa and Ksp values used above; specifically, we must account for:
T ≠ 25 °C, concentrations ≠ 1 M, and salinity ≠ 0.
Taking the last item first, equilibria are often shifted by ionic
strength, as accounted for in the Debye−Huckel equation.
Therefore, the first thing we need to do is characterize equilibria
in seawater, as opposed to dilute aqueous solution. The
recommended nomenclature is to modify the Keq parameter
with an asterisk, K*. Thus, for a typical reaction we have
Figure 1. Scheme showing how reactions 1−3 combine to give CO2-
induced dissociation of CaCO3(s). From Hess’s law, ΔGnet = ΔGa1 −
ΔGa2 + ΔGsp. [C]ceq [D]deq
a A + b B F cC + d D K* ≡
[A]aeq [B]beq (5)
concentration of aqueous protons and lowering the pH; reaction
1 is characterized by Ka1(H2CO3) and ΔGa1°. where [x]eq is the molar concentration in seawater of solute x at
K a1
equilibrium. The relationship between the molar concentration-
CO2 (g) + H 2O(l) F H 2CO3(aq) HooI H+(aq) + HCO3−(aq) based K* and the true, activity-based Keq can be calculated from
{aka, CO2 (aq)} the activity coefficient ratio, as specified in a later section of this
(1) paper (e.g., eq 10).
For reaction 2, the acid ionization of bicarbonate to carbonate Carbonic acid seawater pKa* values can be found in the
and H+ is characterized by Ka2(HCO3−) and ΔGa2°: literature, e.g., pKa1*(H2CO3) = 5.856 and pKa2*(HCO3−) =
K a2
8.925 (e.g., refs 25 and 28). For calcium carbonate, Ksp* can be
HCO3−(aq) HooI H+(aq) + CO32 −(aq) (2) calculated from the saturating concentrations of Ca2+ and CO32−
in seawater:21 10.(49) mM and 62.(5) μM (for aragonite),
Reaction 3, the dissociation of calcium carbonate salt in water, is respectively, giving Ksp* = 0.01049 × 62.5(10−6) = 6.6(10−7).
characterized by Ksp(CaCO3) and ΔGsp°. Standard Gibbs free (The effect of ionic strength on these three equilibrium
energy changes for the three reactions can be calculated from constants is discussed below in the Applications for Physical
ΔG° = −RT ln Keq. Chemistry Students section. It is worth pointing out that while
+H 2 O 10.5 mM is the actual seawater concentration of calcium, 62.5
CaCO3(s) HoooooI Ca 2 +(aq) + CO32 −(aq) μM is the measured saturating concentration of carbonate, above
K sp (3)
which CaCO3 precipitates; it is not the actual seawater [CO32−].
To build calcareous shells, marine organisms must combine Note that [Ca2+] > [CO32−] because the calcium concentration
Ca2+ and CO32− to precipitate CaCO3(s) (reverse of eq 3). is kept high by biological and physical processes.) Using these
However, adding CO2 to seawater produces H+ (eq 1), which in seawater K* values and ΔG* = −RT ln K*, we can calculate that
turn protonates carbonate (reverse of eq 2), lowering its at 25 °C and all concentrations 1 M, ΔGnet* = 33.4(2) − 50.9(4)
concentration; this decrease in carbonate concentration makes + 35.(29) = +17.(8) kJ/mol. Thus, calcium carbonate
calcium carbonate dissociation (eq 3) more spontaneous, thus precipitation (i.e., shell-building) in seawater is still sponta-
making shell-building more difficult. The net reaction for CO2- neous.
induced dissociation of CaCO3 is given by eq 4. From this The marine concentration of calcium, 0.41 g/kg soln (i.e., 10.
equation, we can see that adding CO2 forces the net reaction (23) millimolal), is fairly uniform and homogeneous.21,22
forward, dissolving calcium carbonate shells. However, the saturating concentration of carbonate varies
slightly with temperature and with latitude: For the aragonite
CO2 (g) + H 2O(l) + CaCO3(s) F H 2CO3(aq) + CaCO3(s)
{CO2(aq)} form it is 61 μmol/kg soln in tropical oceans (30 °C) vs 66
μmol/kg soln in the polar Southern Ocean (1 °C).21 This means
F Ca (aq) + 2HCO3−(aq)
2+
(4)
that Ksp(aragonite) is slightly different in the two oceans.
A scheme showing how reactions 1−3 combine to give eq 4 is Before we use these values to calculate Ksp*, a word is in order
shown in Figure 1. Applying Hess’s law, we see that ΔGnet for the concerning concentration units. As is common in solution
2021 https://doi.org/10.1021/acs.jchemed.2c00149
J. Chem. Educ. 2022, 99, 2020−2025
Journal of Chemical Education pubs.acs.org/jchemeduc Article

Table 1. Thermodynamic calculations of Qsp, Ω, and ΔGsp for CaCO3, given carbonate concentrations (and pH) from Orr et
al.21,a
CO2,atmosph (ppm) pHb [CO32−]b (μM) Qsp (× 106) Qsp/Ksp* (= Ω) ΔGspc (kJ/mol)
Tropical Ocean (T = 30 °C)
1800 280 8.12 ± 0.05 275 ± 15 2.89 4.41 3.7(4)
1994 365 8.05 ± 0.02 246 ± 3 2.58 3.93 3.4(5)
2100 788 7.85 ± 0.05 153 ± 14 1.60 2.44 2.2(5)
Polar Southern Ocean (T = 1 °C)
1800 280 8.20 ± 0.05 126 ± 10 1.32 1.86 1.4(2)
1994 365 8.10 ± 0.02 108 ± 2 1.13 1.59 1.0(6)
2100 788 7.85 ± 0.05 56 ± 5 0.59 0.83 −0.4(2)
a
[Ca2+] = 10.(49) mM. bCarbonate concentrations and pH are measured for 1994; they are predicted from models using known [CO2] for the
year 1800 and using predicted [CO2] for 2100.21 cΔGsp = RT ln(Qsp/Ksp*); Ksp* = 6.5(6) × 10−7 and 7.0(9) × 10−7 in the tropical and polar
oceans, respectively. With the inclusion of the effects of global warming, ΔGsp is predicted to be slightly higher in 2100, but only by ≤0.3 kJ/mol.30

chemistry, molality is generally the unit of choice in marine The first thing to note in these data is that, as atmospheric
chemistry, or, more specifically, μmol/kg soln. Because K* carbon dioxide concentrations rise with time, carbonate
values are generally molarity-based, we must use the density of concentrations fall, by ≈12% from 1800 to 1994, and by
seawater, 1.025 kg/L, to convert from molality to molarity: ≈40% from 1994 to 2100. The pH decline (due to the
μmol/kg soln × 1.025 kg/L = μmol/L = μM. So, in seawater, acidification from carbon dioxide hydration and carbonic acid
[Ca2+] = 10.(49) mM and [CO32−] = 62.(5) μM (tropical) or ionization) from 8.2 in 1800 to 8.1 in 1994, and 7.85 in 2100,
67.(65) μM (polar). Thus, Ksp*(CaCO3,aragonite) varies corresponds to proton concentration increases of ≈25% (1800
slightly in the two oceans: 6.5(6) × 10−7 (tropical) vs to 1994) and ≈70% (1994 to 2100). Increasing [H+] favors
7.0(9) × 10−7 (polar). protonation of carbonate to bicarbonate, thus explaining the
Standard conditions (1 M) are only of general interest to decline in [CO32−]. When we compare the two centuries from
chemists; what we really need to know is how this reaction will 1800 to 1994 with the single century from 1994 to 2100, we can
behave under actual ocean conditions. For this calculation, we see that the problem is clearly accelerating,
need to use the reaction quotient, Q, to calculate ΔG under Also of note is the regional disparity. The tropical ocean not
nonstandard state conditions:29 only has a higher temperature and carbonate concentration
(discussed below) but also has a slightly lower pH. The ocean is
ΔG = ΔG◦ + RT ln(Q ) = RT ln(Q /Keq) (6) most decidedly a poorly stirred inhomogeneous solution; its
mixing time scale is on the order of thousands of years.28 Also,
As long as Q > Keq (i.e., Q/Keq > 1), ΔG will be positive, and due to the fast cycling of changes in sunshine, currents and
the reaction will be endergonic. upwelling, and biological activity, many processes are main-
For the moment, let us consider only the dissociation of tained out of equilibrium. Needless to say, predictions based on
calcium carbonate (eq 3): equilibrium thermodynamic parameters (e.g., Keq, ΔG°) may
not be accurate for nonequilibrium systems.
Q sp = [Ca 2 +][CO32 −] (7) Nevertheless, two key findings from Orr et al.21 follow: (1)
The carbonate concentration in the tropical ocean remains
From eq 6, we see that, as long as the carbonate concentration above its saturation limit of 63 μM, but (2) by 2050−2070 in the
exceeds the saturating concentration, Qsp > Ksp*, ΔGsp will be polar Southern Ocean, carbonate will decline to its saturation
positive, and calcium carbonate dissociation will be non- limit of 68 μM. At a predicted carbonate concentration of 56 μM
spontaneous; i.e., precipitation and shell-building are sponta- in 2100, Qsp/Ksp < 1, ΔGsp < 0, and calcium carbonate
neous. (In the literature, the ratio Qsp/Ksp* is often referred to as dissolution will be spontaneous.17,31 Of course, the fact that
Ωi, the degree of saturation of seawater with respect to salt i: If Ωi calcium carbonate shells dissolve spontaneously in solutions
> 1, seawater is supersaturated, and the precipitation of salt i is undersaturated with carbonate does not tell us how quickly this
spontaneous; undersaturated seawater has Ωi < 1, and will occur, nor whether living calcifying organisms might be
dissociation is spontaneous.) Orr et al. reported that, in 1994, harmed by this process. In fact, shells begin to dissolve within
carbonate concentrations averaged 108 ± 2 μM in the Southern days of solutions falling below the carbonate saturation
Ocean (T = 1 °C, 70° S latitude) vs 246 ± 3 μM in tropical concentration,32,33 even among living organisms.22 Buth has
waters (T = 30 °C).21 The underlying reasons for this dramatic published in this Journal an excellent upper division under-
difference in carbonate concentration are discussed below. The graduate laboratory project characterizing the effect of seawater-
resulting values of Qsp, Ω, and ΔGsp for calcium carbonate carbonation on pH, sea shell mass, and concentrations of
dissociation are presented in Table 1. bicarbonate and calcium.5
Before we discuss the data in Table 1, a word is in order Orr et al.21 exposed live pteropods, typical calcifying plankton
regarding their sources. All 1994 data were measured in the Southern Ocean, to carbonate-undersaturated solutions
experimentally.21 Atmospheric [CO2(g)] in 1800 was measured similar to those predicted for the Southern Ocean in 2100.
from ice core samples, and the year 2100 value was predicted Within 48 h, etched pits formed on the shell surfaces, and
using the “business-as-usual” scenario IS92a from the United marked dissolution was observed at the growing edges of the
Nations IPCC.21 All other tabulated values for 1800 and 2100 shell. Similarly, Kuffner et al.23 reported that, at the acidified pH
were calculated from [CO2(g)]atm using standard marine CO2/ of 7.85, the growth of calcifying algae was severely inhibited. In
CO32− and alkalinity equations.21,28 fact, Feely et al.22 reviewed extensive evidence showing that,
2022 https://doi.org/10.1021/acs.jchemed.2c00149
J. Chem. Educ. 2022, 99, 2020−2025
Journal of Chemical Education pubs.acs.org/jchemeduc Article

even when the carbonate concentration is maintained at Table 2. Activity coefficients for seawater ions,a calculated
supersaturating levels (i.e., CaCO3 precipitation is sponta- from the Davies Equation (A = 0.509 at 25 °C in water, ε =
neous), decreasing [CO32−] lowers the calcification rate of 78.4), and measured28
almost all calcifying organisms studied, “from single-celled
ion γi,calc γi,meas
protists to reef-building corals”. (Although Fabry cast some
doubt on this conclusion,34 Beaufort et al. have confirmed it and Na+ 0.69 0.668
explained the discrepancies.35) Cl− 0.69 0.666
Orr et al.21 concluded that if a calcifying organism “cannot HCO3− 0.69 0.570
grow its protective shell, we would not expect it to survive in H+ 0.69 0.590
waters that become undersaturated with respect to aragonite.” OH− 0.69 0.255
Pteropods are an important part of the diet of carnivorous H2PO4− 0.69 0.453
zooplankton, as well as larger cold-water fish and mammals Mg2+ 0.23 0.240
including salmon, mackerel, herring, cod, and baleen whales. Ca2+ 0.23 0.203
Loss of these calcifying organisms could thus have a significant SO42− 0.23 0.104
deleterious effect on the entire marine ecosystem of the CO32− 0.23 0.039
Southern Ocean by the end of this century. Orr et al.21 HPO42− 0.23 0.043
a
predicted that the problem will spread to higher latitude oceans Thermodynamic objections have been raised (and countered37)
by 2150−2200. In fact, there is evidence that seasonal decreases regarding the calculation and determination of activity coefficients for
in carbonate concentration have already impacted calcifying individual ions, as opposed to the mean activity coefficient of an
organisms in the midlatitude from the Pacific coast of Canada to electrolyte. Nevertheless, activity coefficients of individual ions are
Mexico18 and in the Baltic and Black Seas.36 commonly used in marine chemistry and many other subdisciplines of
chemistry.
Applications for Physical Chemistry Students
The calculations above are appropriate for general chemistry. a H+aHCO−3 γH+γHCO− [H+][HCO−3 ]
3
Additionally, there are two interesting aspects of this system that K a1 = = ×
aCO2 γCO [CO2 ]
are appropriate for physical chemistry students, namely, the 2

effects of salinity and temperature on equilibrium constants. γH+γHCO−


3
= × K a1*
Salinity and Ionic Strength γCO (10)
2
We noted above that the Ka* values for carbonic acid and the
Ksp* of calcium carbonate were 1−2 orders of magnitude greater Thus

jij γH+γHCO−3 zyz


in seawater than the activity-based equilibrium constants

pK a1* = pK a1 + logjjj zz
zz
jj γ z
measured in dilute aqueous solution. These differences lowered

k {
ΔGnet for the CO2-induced dissociation of calcium carbonate by
30% (from 25 to 18 kJ/mol). At least part of the explanation for CO2

these dramatic increases in reaction spontaneity is due to ionic


strength and, thus, can be explained using the limiting Debye− Measured in dilute aqueous solution (where all γi ≈ 1, so K* ≈
Huckel law, which predicts the decrease in the activity Keq), the pKa1 for carbonic acid is 6.365. However, the pKa1* for
coefficient of an ion (γi) with increasing ionic strength (for I < carbonic acid in seawater is 5.856; the calculation below shows
0.01): that the decrease in ion activity in seawater explains the increase
in Ka* (i.e., decrease in pKa*). Using eq 9 and measured activity
log(γi) ≈ −Azi 2 I (8) coefficients (Table 2), we get
−3/2
where A is a collection of constants, including the term (T) , pK a1* ≈ 6.365 + log(0.590 × 0.570/1) = 6.365 − 0.473 = 5.89
1
zi is the charge number of ion i, and I (ionic strength) ≡ 2 ∑i cizi2 , (11)
where ci is the concentration of ion i, and the sum is carried out (Here we assume that H2CO3, a neutral solute, has γ ≈ 1 in
for all ions present in solution. seawater.)
For seawater, I ≈ 0.7,28 so the Debye−Huckel limiting law Similarly, for carbonic acid’s second ionization,

ij γ + × γ 2− yz
does not hold. Instead, the Davies equation,28 which is accurate
j H CO3 z
pK a2* = pK a2 + logjjj zz
jj γ − zzz
for I < 0.5, comes closer:
i y
log(γi) ≈ −Az i 2jjjj − 0.2 × I zzzz k {
k1 + I {
I HCO3
(9)
= 10.24 + log(0.590 × 0.039/0.570)
Activity coefficients have been measured experimentally for
the important ions in seawater (Table 2). We can see that, for = 10.32 − log(0.0404)
some anions (e.g., OH−, CO32−, HPO42−), the measured activity = 8.93 (12)
coefficients in seawater are substantially lower than the value
predicted by the Davies equation. These anions form especially The agreement with the measured seawater value of pKa2* =
strong complexes and ion pairs that lower their activity below 8.925 is perfect.
that predicted by the Davies equation, which accounts only for Finally, for the solubility product of calcium carbonate,
long-range electrostatic interactions. measured in dilute aqueous solution, Ksp = 4.5(10−9), whereas in
For carbonic acid’s first ionization, the thermodynamic seawater, Ksp* = 6.85(10−7). Using eq 9 and measured activity
(activity-based) equilibrium constant is coefficients in seawater (Table 2), we get
2023 https://doi.org/10.1021/acs.jchemed.2c00149
J. Chem. Educ. 2022, 99, 2020−2025
Journal of Chemical Education pubs.acs.org/jchemeduc Article

pK sp* = pK sp + log(γCa 2+ × γCO2−) CO2 (aq) + H 2O(l) + CO32 −(aq) F 2HCO3−(aq) (15)
3

= 8.35 + log(0.203 × 0.039) = 6.25 (13) This reaction is spontaneous in seawater under standard
conditions (all concentrations 1 M: ΔG°net = 33.4(2) − 50.9(4)
Again, the agreement with the measured seawater value of = −17.5 kJ/mol) and is essentially at equilibrium under ocean
pKsp* = 6.17 is excellent. Thus, the fact that K* in seawater is conditions (ΔG = +0.1 kJ/mol, see Supporting Information).
higher than Keq in dilute aqueous solution is explained by the Thus, raising [CO2(aq)] forces the reaction forward, which in
lower activity coefficients of solute ions in seawater. Interested turn raises the bicarbonate concentration and lowers the
readers are referred to Zeebe and Wolf-Gladrow for an expanded carbonate concentration in the polar Southern Ocean. This
discussion and further details.28 reaction could also explain why the Southern Ocean is slightly
Temperature and Heat Capacity more alkaline: pH (Southern Ocean) = 8.10 > 8.05 in the
tropical ocean (Table 1). Note that the product of eq 15 is
We noted above that the carbonate concentration in the tropical bicarbonate, which is a weak base.
ocean is currently more than double that in the polar Southern Although higher [CO2(aq)] partially explains the lower
Ocean: 246 vs 108 μM. There are at least three possible carbonate concentration in the polar Southern Ocean,
explanations for this disparity: (1) Carbonate-producing explanation 3 is also significant: Carbonate concentrations are
reactions are endothermic; (2) [CO2(aq)] is higher in the maintained away from equilibrium due to biological and physical
Southern Ocean (due to its higher solubility at lower processes (e.g., photosynthesis, upwelling and currents, etc.)
temperature, and to deep water upwelling and organic matter that differ dramatically in the two oceans.38
remineralization21); or (3) the carbonate concentration is
maintained away from equilibrium due to biological and physical
processes (e.g., photosynthesis, upwelling and currents, etc.).
Explanation 1. It turns out that both ionization reactions of
■ SUMMARY
Rising atmospheric carbon dioxide is expected to acidify the
carbonic acid are indeed endothermic. From the Gibbs− ocean, resulting in a decline in carbonate concentration. A
Helmholtz equation (eq 14), we see that ln Keq for endothermic number of complex models have been used to predict how future
reactions decreases with 1/T (i.e., increases with T): scenarios of carbon dioxide increase will affect ocean acidity and
carbonate.21,22 For example in this Journal, Weston’s “Climate
∂ln Keq −ΔH ° Change and Its Effect on Coral Reefs”24 included over 20
=
1 equations, while Bozlee et al’s “Simplified Model to Predict the
∂ () T
R
(14) Effect of Increasing Atmospheric CO2 on Carbonate Chemistry
in the Ocean” included 19 equations, of which five were
Although the increase in Ka with temperature would explain
simultaneous equations that were distilled down to one using the
the higher carbonate concentration in tropical oceans, one must
quadratic equation.25 In contrast, here we have used carbonate
keep in mind that a negative heat capacity change for the
concentrations predicted by models, along with a simple set of
reaction (ΔC°P) causes ΔH° to become negative at higher T.
equilibrium thermodynamic calculations based only on the
Although ΔCP is negative for carbonic acid ionization (ΔC°P =
acidity of carbonic acid and the solubility of calcium carbonate.
−378 and −17.4 J/mol/K for the first and second carbonic acid
These equilibria allow first-year chemistry students to calculate
ionization reactions, respectively; see Table S1 and Figure S1,
for themselves the spontaneity of calcium carbonate shells
Supporting Information), ΔH°a is still positive over the entire
dissolving in the polar Southern Ocean by the year 2100. The
1−38 °C temperature range, and Ka at 38 °C is nearly twice its
calculations we present here bring critical real-world environ-
value at 1 °C, for both reactions (Figure S2, Table S1). Although
mental chemistry into the introductory and physical chemistry
it looks as if this might explain the higher carbonate
classrooms. Student responses to the case study and the problem
concentration in the tropical ocean, note that the net Keq of
sets (Supporting Information) have been quite positive, most
both reactions combined is Ka1Ka2 = 3(10−17) at 30 °C; while
appreciating the application of “dry” thermodynamic theory to a
this is three times the value at 1 °C, it is still so small as to be an
pressing real-world problem.


insignificant contributor of aqueous carbonate.
Furthermore, the other carbonate-producing reaction in
seawater, CaCO3(s) dissociation, is slightly exothermic, with ASSOCIATED CONTENT
ΔH°298 = −1.9 kJ/mol (Figure S3), and also quite endergonic *
sı Supporting Information
(Ksp ≈ 10−7). Thus, thermodynamics cannot explain the
difference in steady state carbonate concentrations between The Supporting Information is available at https://pubs.ac-
tropical and polar oceans. s.org/doi/10.1021/acs.jchemed.2c00149.
Explanation 2. The surface water in polar oceans has higher Carbonic acid dissociation thermodynamic parameters, T
[CO2(aq)] than tropical oceans for two reasons: (i) Like most = 0 to 38 °C; plot of temperature dependence of carbonic
gases, aqueous solubility of CO2 is higher at lower temperature, acid dissociation enthalpies; plot of temperature depend-
and (ii) the upwelling of deep water is much more substantial in ence of carbonic acid Ka1 and Ka2; plot of temperature
the polar Southern Ocean.21 Deep ocean water is known to have dependence of the solubility (ΔG° sp ) of CaCO 3
a higher [CO2(aq)] than surface waters due to the biological (aragonite); and calculation of ΔG for the seawater
process of remineralization that is carried out by organisms in reaction of aqueous CO2 + carbonate (PDF, DOCX)
the ocean floor sediment.21,28 (Remineralization is simply the
oxidative catabolism of organic matter to CO2 + H2O + simple General chemistry/first-year student worksheet (PDF,
nutrient molecules.) Furthermore, [CO2(aq)] and [CO32−] are DOCX)
reciprocally related,25 due to the protonation of carbonate by
carbonic acid: Physical chemistry student worksheet (PDF, DOCX)

2024 https://doi.org/10.1021/acs.jchemed.2c00149
J. Chem. Educ. 2022, 99, 2020−2025
Journal of Chemical Education


pubs.acs.org/jchemeduc Article

AUTHOR INFORMATION (18) Feely, R. A.; Sabine, C. L.; Hernandez-Ayon, J. M.; Ianson, D.;
Hales, B. Evidence for Upwelling of Corrosive″ Acidified″ Water onto
Corresponding Author the Continental Shelf. Science 2008, 320 (5882), 1490−1492.
Todd P. Silverstein − Department of Chemistry, Willamette (19) Zeebe, R. E. History of Seawater Carbonate Chemistry,
University, Salem, Oregon 97301, United States; orcid.org/ Atmospheric CO2, and Ocean Acidification. Annu. Rev. Earth Planet.
Sci. 2012, 40, 141−165.
0000-0003-3349-2876; Email: tsilvers@willamette.edu (20) Zeebe (2012, ref 19) reported that the decrease in ocean pH due
Complete contact information is available at: to anthropogenic CO2 release will be about five times larger and 100
https://pubs.acs.org/10.1021/acs.jchemed.2c00149 times faster than during a typical deglaciation period.
(21) Orr, J. C.; Fabry, V. J.; Aumont, O.; Bopp, L.; Doney, S. C.; Feely,
R. A.; Gnanadesikan, A.; Gruber, N.; Ishida, A.; Joos, F.; et al.
Notes
Anthropogenic Ocean Acidification over the Twenty-First Century and
The author declares no competing financial interest. Its Impact on Calcifying Organisms. Nature 2005, 437 (7059), 681−


686.
(22) Feely, R. A.; Sabine, C. L.; Lee, K.; Berelson, W.; Kleypas, J.;
ACKNOWLEDGMENTS Fabry, V. J.; Millero, F. J. Impact of Anthropogenic CO2 on the CaCO3
The invaluable insights and suggestions of my colleague J. System in the Oceans. Science 2004, 305 (5682), 362−366.
Charles (Chuck) Williamson are duly noted here. His careful (23) Kuffner, I. B.; Andersson, A. J.; Jokiel, P. L.; Rodgers, K. S.;
reading and thoughtful analysis are greatly appreciated. Mackenzie, F. T. Decreased Abundance of Crustose Coralline Algae


Due to Ocean Acidification. Nat. Geosci. 2008, 1 (2), 114−117.
(24) Weston, R. E., Jr Climate Change and Its Effect on Coral Reefs. J.
REFERENCES Chem. Educ. 2000, 77 (12), 1574−1577.
(1) Bayer, R.; Hudson, B.; Schneider, J. Transformation of Chemistry (25) Bozlee, B. J.; Janebo, M.; Jahn, G. A Simplified Model to Predict
Experiments into Real World Contexts. J. Chem. Educ. 1993, 70 (4), the Effect of Increasing Atmospheric CO2 on Carbonate Chemistry in
323−324. the Ocean. J. Chem. Educ. 2008, 85 (2), 213−217.
(2) Kerber, R. C.; Akhtar, M. J. Getting Real: A General Chemistry (26) Fasman, G. D. Handbook of Biochemistry and Molecular Biology,
Laboratory Program Focusing on″ Real World″ Substances. J. Chem. 3rd ed.; CRC: Boca Raton, FL, 1976; Vol. 4.
Educ. 1996, 73 (11), 1023−1025. (27) Brown, T. L.; LeMay, H. E., Jr.; Bursten, B. E.; Murphy, C. J.;
(3) Hostettler, J. D. Introduction to the″ Real World″ Examples Woodward, P. K. Chemistry: The Central Science, 12th ed.; Prentice
Symposium. J. Chem. Educ. 1983, 60, 1031−1032. Hall: Boston, 2012.
(4) Kozlowski, A. W. Using Real World Examples in a Laboratory (28) Zeebe, R. E.; Wolf-Gladrow, D. CO2 in Seawater: Equilibrium,
Program. J. Chem. Educ. 1983, 60, 1039−1040. Kinetics, Isotopes; Elsevier: Amsterdam, 2001.
(5) Buth, J. M. Ocean Acidification: Investigation and Presentation of (29) Silverstein, T. P. The Reaction Quotient Is Useful after All. J.
the Effects of Elevated Carbon Dioxide Levels on Seawater Chemistry Chem. Educ. 2005, 82 (8), 1149−1149.
and Calcareous Organisms. J. Chem. Educ. 2016, 93 (4), 718−721. (30) Cao, L.; Caldeira, K.; Jain, A. K. Effects of Carbon Dioxide and
(6) Breslin, V. T.; Sañudo-Wilhelmy, S. A. The Lead Project. An Climate Change on Ocean Acidification and Carbonate Mineral
Environmental Instrumental Analysis Case Study. J. Chem. Educ. 2001, Saturation. Geophys. Res. Lett. 2007, 34 (5), L05607.
78 (12), 1647−1651. (31) Khatiwala et al. (2009, ref 17) reported that the Southern Ocean
(7) Brzenk, R.; Moore, A.; Alfano, M. J.; Buckley, P. T.; Newman, M. is the primary conduit for human-generated carbon dioxide entering the
E.; Dunnivant, F. M. Understanding the Greenhouse Effect: Is Global ocean.
Warming Real? An Integrated Lab-Lecture Case Study for Non-Science (32) Betzer, P. R.; Byrne, R. H.; Acker, J. G.; Lewis, C. S.; Jolley, R. R.;
Majors. J. Chem. Educ. 2000, 77 (12), 1602−1603. Feely, R. A. The Oceanic Carbonate System: A Reassessment of
(8) Doherty, M. P. A Lively and Surprising Toxicology Case Study. J. Biogenic Controls. Science 1984, 226 (4678), 1074−1077.
Chem. Educ. 1994, 71 (10), 860. (33) Byrne, R. H.; Acker, J. G.; Betzer, P. R.; Feely, R. A.; Cates, M. H.
(9) Tro, N. J. Chemistry as General Education. J. Chem. Educ. 2004, 81 Water Column Dissolution of Aragonite in the Pacific Ocean. Nature
(1), 54−57. 1984, 312 (5992), 321−326.
(10) Jones, M. B.; Miller, C. R. Chemistry in the Real World. J. Chem. (34) Fabry, V. J. Marine Calcifiers in a High-CO2 Ocean. Science
Educ. 2001, 78 (4), 484−487. 2008, 320 (5879), 1020−1022.
(11) Brink, C. P.; Goodney, D. E.; Hudak, N. J.; Silverstein, T. P. A (35) Beaufort, L.; Probert, I.; de Garidel-Thoron, T.; Bendif, E. M.;
Ruiz-Pino, D.; Metzl, N.; Goyet, C.; Buchet, N.; Coupel, P.; Grelaud,
Novel Spiral Approach to Introductory Chemistry Using Case Studies
M.; et al. Sensitivity of Coccolithophores to Carbonate Chemistry and
of Chemistry in the Real World. J. Chem. Educ. 1995, 72 (6), 530−532.
Ocean Acidification. Nature 2011, 476 (7358), 80−83.
(12) Silverstein, T. P.; Goodney, D. E.; Holman, K. L. M.; Kirk, S. R.;
(36) Tyrrell, T.; Schneider, B.; Charalampopoulou, A.; Riebesell, U.
Willemsen, J. J.; Williamson, J. C. Model Courses. Chem. Eng. News
Coccolithophores and Calcite Saturation State in the Baltic and Black
2003, 81 (7), 6.
Seas. Biogeosci. 2008, 5 (2), 485−494.
(13) Silverstein, T.; Hudak, N. J. Where Have All the Chem Majors
(37) Fraenkel, D. Single-Ion Activity: Experiment versus Theory. J.
Gone? CUR Quart. 1994, 14, 127−130.
Phys. Chem. B 2012, 116 (11), 3603−3612.
(14) Middlecamp, C. H.; Jordan, T.; Shachter, A. M.; Kashmanian
(38) Orr, J. C. Ocean Acidification and Calcification (personal
Oates, K.; Lottridge, S. Chemistry, Society, and Civic Engagement
communication), 2006.
(Part 1): The SENCER Project. J. Chem. Educ. 2006, 83 (9), 1301−
1307.
(15) Middlecamp, C. H.; Phillips, M. F.; Bentley, A. K.; Baldwin, O.
Chemistry, Society, and Civic Engagement (Part 2): Uranium and
American Indians. J. Chem. Educ. 2006, 83 (9), 1308−1312.
(16) Kirk, S. R.; Silverstein, T. P.; Willemsen, J. J. Teaching
Biologically Relevant Chemistry throughout the Four-Year Chemistry
Curriculum. J. Chem. Educ. 2006, 83 (8), 1171−1175.
(17) Khatiwala, S.; Primeau, F.; Hall, T. Reconstruction of the History
of Anthropogenic CO 2 Concentrations in the Ocean. Nature 2009,
462 (7271), 346−349.

2025 https://doi.org/10.1021/acs.jchemed.2c00149
J. Chem. Educ. 2022, 99, 2020−2025

You might also like