AAMC IQB Chemistry [Ext]

 

Consider the reaction rate described by the following equation.
Rate = k[A]m[B]n
The overall reaction is said to be second order when:
This question asks us to determine under which conditions the overall reaction rate is of the second order. Answering this question requires us to delve into the theory of reaction rates and orders.
The rate law describes how reactant concentrations influence the rate of a reaction. The general form is:
Rate = k[A]m[B]n
Here:
– “Rate” signifies the reaction rate.
– “k” denotes the rate constant specific to the reaction at a set temperature.
– “[A]” and “[B]” represent the concentrations of reactants A and B.
– “m” and “n” are reaction orders concerning A and B. These values are typically 0 (zeroth order), 1 (first order), 2 (second order), etc., determined experimentally. The overall order is the sum of the individual orders or m + n.

In this option, m and n both equal 0, which would make the overall order of the reaction 0 (m + n = 0 + 0 = 0). Thus, this choice is incorrect because a zeroth-order reaction does not describe a second-order reaction.
For Option B, where m=1 and n=1, the rate equation becomes: Rate = k[A][B]. The reaction has a first order for both A and B. The rate of this reaction is directly proportional to the concentration of A and directly proportional to that of B. By determining the overall order, which is the sum of individual orders (1 + 1 = 2), the reaction is second order, making this choice correct.
Here, m and n are 2, leading to an overall order of 4 (m + n = 2 + 2 = 4). This isn’t a second-order reaction; instead, it’s fourth-order, so this choice is incorrect.
The value of k (the rate constant) does not affect the order of the reaction. The order depends on the exponents m and n. Since m and n are variable in this option, there’s no concrete way to confirm that the reaction is second order. This choice is, therefore, incorrect.
The question requires us to identify which constant will help determine the amount of Pb2+ that will precipitate when HCl is introduced to a Pb(NO3)2(aq) solution.
Precipitation occurs when solutions are mixed and a solid forms. The solubility product constant, denoted by Ksp, is a value used to describe the solubility of a compound in water. It represents the equilibrium between the solid form of an ionic compound and its ions in solution. The lower the Ksp value, the less soluble the compound is.
Ka for HCl isn’t relevant for this reaction. HCl is a strong acid and fully dissociates in water. Its acid dissociation constant (Ka) would not provide information on the precipitation of PbCl2. Therefore, this option is incorrect.
HNO3 is not directly involved in the precipitation reaction. The Ka value for HNO3 would tell us about the extent to which nitric acid ionizes in solution, which is irrelevant to the precipitation of PbCl2. So, this option is not correct.
For the formation of a precipitate when HCl is added to Pb(NO3)2, the following reaction is expected:
Pb2+(aq) + 2Cl(aq) → PbCl2(s)
Here, PbCl2 will precipitate if the ion product exceeds the Ksp of PbCl2.
Option C is correct because the Ksp for PbCl₂ is necessary to determine the solubility of PbCl₂ in the solution. If the ion product ([Pb2+][Cl]2) exceeds the Ksp of PbCl2, then PbCl2 will precipitate from the solution.
Keq for the reaction Pb2+ + 2 e ⇌ Pb relates to the equilibrium constant for reducing lead ions to metallic lead. This redox reaction does not provide information on the solubility of PbCl2 or its precipitation from the solution. Thus, this option is incorrect.
The question requires us to identify the ground-state electron configuration for nitrogen, which corresponds to its most stable electronic state.

Electrons are arranged in different energy levels (shells) and sublevels (s, p, d, f) within those levels. Each sublevel contains a specific number of orbitals, and each orbital can hold up to two electrons with opposite spins. Beginning with the 1s orbital, electrons occupy progressively higher energy levels, including the 2s and 2p orbitals. As the energy levels rise, subsequent orbitals like the 3s, 3p, and 3d come into play. This process continues with 4s, 4p, 4d, and 4f orbitals in higher energy levels.

The Aufbau Principle is a key concept in chemistry that governs how electrons occupy atomic orbitals within an atom. It states that electrons first enter orbitals of lowest energy before moving to higher energy orbitals. This means that as electrons are added to an atom, they start by occupying the lowest energy level orbitals and only move to higher energy orbitals once the lower ones are filled.

1s22s12p4 suggests nitrogen has 8 electrons, which is incorrect. It also violates Hund’s rule, which states that electrons occupy degenerate orbitals singly before pairing up.
Nitrogen (N) has an atomic number of 7, meaning it has 7 electrons. For this question, we are tasked with finding the ground-state electron configuration for nitrogen (i.e., the electronic configuration where the electrons are in the lowest possible energy levels). Option B, 1s22s22p3, is the ground-state electron configuration for nitrogen because it accurately depicts the distribution of all 7 electrons of nitrogen in the lowest energy orbitals available, adhering to the Aufbau principle.
1s12s22p4 implies that one of the 1s electrons has been excited to a higher energy level, which would not correspond to the most stable state of nitrogen, which would fill the 1s orbital first.
1s22s22p23s1 suggests nitrogen has 8 electrons, and it has an excited electron in the 3s orbital, which also doesn’t correspond to the most stable state of nitrogen, which would fill the 2p orbital first.
The question asks us to identify the factor determining the pH at which a color change occurs when a weak acid is titrated with a base in the presence of an indicator.

An indicator is a versatile tool in chemistry, particularly in acid-base interactions. It functions as a reversible switch between being a weak acid and its corresponding weak base partner. This balance results in the ability of indicators to display color shifts depending on the pH of their environment. Indicators showcase this within a relatively narrow pH range, typically centered around a value known as the indicator’s pKa. This pKa holds significance as it represents the pH at which exactly half of the indicator molecules exist in their acidic form and the remaining half in their basic form.

The final concentration of HA does not determine the pH at which the indicator changes color. The concentration might affect the total volume needed for the titration but not the pH of the color change.
The final concentration of HIn does not determine the exact pH at which the color change occurs. While the indicator’s concentration might affect the color’s intensity, the pH at which the color change happens is primarily determined by its pKa.
The initial concentration of HA affects how much NaOH is needed to reach the equivalence point of the titration but doesn’t determine the pH at which the indicator changes color. The pH of the color change is intrinsic to the indicator itself and its pKa.
The pH at which the indicator changes color is dependent on its pKa. When the pH of the solution is equal to the pKa of the indicator, half of the indicator molecules are protonated (HIn), and half are deprotonated (In), resulting in a visible color change. Therefore, we can conclude that option D is correct.
This question asks us to determine the moles of Al3+ reduced when a certain charge (0.1 faraday) is passed through an electrochemical cell.

A faraday is the charge carried by one mole of electrons. When considering metal deposition from its ions during electrolysis, the number of moles of a metal deposited (or ions reduced) is directly proportional to the number of faradays of electricity passed.

For the reduction of Al3+ to aluminum metal, the half-reaction is:
Al3+ + 3e → Al

This indicates that 3 moles of electrons (or 3 faradays) are required to reduce 1 mole of Al3+.

Given the 3:1 stoichiometry between electrons and Al3+ ions, 0.1 faraday would reduce:
Moles of Al3+ = (0.1 faradays) / 3 = 0.033 moles of aluminum

Therefore, we can conclude that 0.033 moles of Al3+ are reduced by 0.1 faradays.

This answer suggests that 0.1 faraday of charge could reduce half of 0.1 moles of a divalent ion (like Ca2+). This doesn’t apply to Al3+, which is trivalent.
This answer is not easily derived from clear calculations involving the trivalent nature of Al3+ and the given 0.1 faraday of charge.
This answer implies that 0.1 faraday of charge can reduce 0.1 moles of a monovalent ion (like Na+). This isn’t correct for Al3+, which requires 3 moles of electrons for every mole of ion reduced.
This question asks us to determine which species among the given options has the highest electron affinity. This requires us to recall our understanding of

Electron affinity is the energy change that occurs when an electron is added to a neutral atom. A higher (more negative) electron affinity indicates a greater attraction for an added electron. Typically, non-metals have higher electron affinities than metals.

Cd(s) is a metal; typically, metals have low electron affinities because they tend to lose electrons rather than gain them. This is evident from the reaction where Cd is oxidized, meaning it loses electrons. Therefore, this choice is not correct.
From the given reaction, Cd(s) loses 2 electrons to form Cd2+(aq), meaning Cd is being oxidized. On the other hand, 2H+(aq) gains these 2 electrons to form H2(g), which means H+(aq) is being reduced.

The fact that H+(aq) accepts electrons more readily in this reaction suggests that it has a higher electron affinity compared to the other species listed.

Therefore, H+(aq) (option B) has the highest electron affinity among the choices.

Cd2+(aq) is an ion that has already lost electrons. It’s unlikely to have a high electron affinity since it’s in a cationic state. This choice is not the best representation of high electron affinity.
H2(g) is a neutral molecule and doesn’t typically have a strong desire to gain additional electrons, especially when compared to H+(aq), which is actively seeking electrons to become neutral. This makes this choice less likely to have the highest electron affinity compared to H+(aq).
This question asks us to determine the primary reason for the difference in boiling points between ammonia (NH₃) and phosphine (PH₃). This prompts us to recall how intermolecular forces affect the physical characteristics of compounds.
The boiling point of a substance is largely influenced by the types and strengths of the forces acting between its molecules. Intermolecular forces (IMFs) are the forces that act between molecules, and they determine various physical properties of substances, including boiling points.

Ammonia (NH₃) is known to form hydrogen bonds, which are particularly strong dipole-dipole attractions that arise when a hydrogen atom bonded to a highly electronegative atom (nitrogen, oxygen, or fluorine) is attracted to another electronegative atom in a nearby molecule.

The basicity of a molecule doesn’t directly correlate with its boiling point. While it’s true that ammonia is a stronger base than phosphine due to the higher electronegativity of nitrogen compared to phosphorus, this difference in basicity doesn’t account for the boiling point discrepancy.
While bond strengths can play a role in certain properties, the boiling point is determined by intermolecular forces, not bond strength within the molecule (intramolecular forces). Therefore, we can conclude that this statement does not directly influence the boiling point difference between ammonia and phosphine.
This statement is generally misleading. Higher molecular weight compounds often have higher boiling points due to increased London dispersion forces (weak attractions between molecules due to temporary changes in their electron distribution). However, NH₃ and PH₃ are not significantly different in molecular weight, and the primary factor affecting their boiling points is the presence or absence of hydrogen bonding.

In the case of ammonia, the hydrogen atoms of one molecule can form hydrogen bonds with the nitrogen atom of another molecule, resulting in a network of strong hydrogen bonds that increase the boiling point. On the other hand, phosphine (PH₃) does not form significant hydrogen bonds. Phosphorus is less electronegative than nitrogen, leading to a weaker dipole in PH₃ compared to NH₃. Without the strong hydrogen bonding in ammonia, phosphine has weaker intermolecular attractions and, thus, a lower boiling point.

Therefore, we can conclude that the statement “ammonia forms stronger intermolecular hydrogen bonds than phosphine” best explains the difference in the boiling points of NH₃ and PH₃, making option D the correct answer.

This question asks us to determine how the volume of a gas will change when its pressure is increased, keeping the temperature constant. This prompts us to consider both the behavior of ideal gases and the characteristics of real gases, especially at high pressures.
Firstly, to calculate the 0.02 L number cited in all of the answer choices, we must use Boyle’s Law for an ideal gas (at constant temperature).

This is the volume for an ideal gas. Real gases deviate from ideal behavior, especially at high pressures.
The equation (P₁V₁ = P₂V₂) relates initial and final pressure and volume values, where P₁ represents the starting pressure (1 atm), V₁ is the initial volume (10 L), P₂ is the ending pressure (500 atm), and V₂ is the volume we’re trying to find. Using Boyle’s Law for ideal gases, we rearrange the equation to find V₂ = (P₁ x V₁) / P₂. Substituting P₁ = 1 atm, V₁ = 10 L, and P₂ = 500 atm, we get V₂ = (1 x 10) / 500 = 0.02 L.

However, in the ideal gas model, two primary assumptions are made:
1) Gas molecules themselves occupy no volume.
2) There are no intermolecular forces (neither attraction nor repulsion) between gas molecules.

These assumptions hold well at low pressures and high temperatures. However, these assumptions break down at very high pressures, like 500 atm. This is because as pressure increases, gas molecules are compressed closer together. As they get closer, the volume they occupy becomes more significant compared to the total volume of the gas. In other words, the “empty space” between molecules in a gas decreases, and the actual volume of the gas molecules becomes more relevant.

Therefore, we can conclude that option B is correct because, at high pressures, the volume occupied by individual gas molecules becomes more significant.

While repulsions between gas molecules might exist, the predominant factor at such high pressures is the actual volume of the gas molecules, not repulsions between them.
The frequency and force of collisions against the container walls define the pressure of the gas, not its volume. The increase in collisions due to pressure doesn’t correlate directly with a change in volume due to non-ideal behavior.
This question asks us to determine how the volume of a gas will change in response to an increase in temperature, given certain conditions. This prompts us to think about ideal gas behavior, the underlying assumptions, and the interplay between pressure, volume, and temperature.

The ideal gas law is represented by the equation: PV = nRT. Where
– P represents pressure.
– V represents volume.
– n is the number of moles of the gas.
– R is the universal gas constant.
– T denotes absolute temperature, measured in Kelvin.

From the ideal gas law, we can derive several relationships between the parameters of gases. One such relationship is between volume (V) and temperature (T), particularly when pressure (P) and the number of moles (n) are constant. This relationship, known as Charles’s Law, states that for a fixed quantity of gas and constant pressure, the volume is directly proportional to its absolute temperature.

Although an increase in temperature usually results in a volume increase, this direct proportionality only remains true if the pressure is constant. The term “always” in this choice can be misleading.
This choice contradicts the principles of Charles’s Law, making it incorrect.
Mathematically, if we consider a constant amount of gas and pressure, the ideal gas law becomes: V = (nR/P) × T. This equation illustrates that if n and P are constant, the volume V will change in direct proportion to the temperature T. It’s essential to note that this relationship holds true only when the temperature is measured in Kelvin, starting from absolute zero (the lowest temperature possible where particles have minimal kinetic energy).

Therefore, we can conclude that option C is correct as the statement correctly reflects the predictions of both the ideal gas law and Charles’s Law. When the temperature rises with constant pressure, the volume will increase, aligning with the derived equation: V = (nR/P) x T.

This choice opposes the predictions of the ideal gas law and Charles’s Law.
This question asks us to determine the key characteristic that makes a material a good insulator but a poor conductor. This involves understanding the fundamental differences between conductors and insulators based on the mobility of their electrons.

Electrical conductivity in materials depends on the ease with which electrons can move within the material. Conductors, such as metals, have a “sea” of delocalized electrons that can move freely, allowing electric current flow. Conversely, insulators are materials that do not allow for easy movement of electrons, thus preventing or significantly reducing the flow of electric current.

Every material contains electrons as they are fundamental particles in atoms. It is important to note that no insulator is perfect, as even insulators have some mobile charges capable of conducting current. Moreover, when a strong enough voltage is applied, insulators can become conductive as the electric field forcibly removes electrons from atoms.
This is not necessarily true for insulators. The magnitude of the electric field inside a material does not solely determine whether it is a conductor or insulator. While insulators resist the flow of electricity, they can still be polarized by an external electric field.
The movement of atoms from one lattice site to another is not a primary factor determining electrical conductivity. This describes ionic conductivity (the ability of ions to move within a material), which differs from electronic conductivity. Thus, this choice is incorrect.
In insulators, the valence electrons (outermost electrons) are tightly bound to their parent atoms. Because of this strong binding, a significant amount of energy is required to detach them and enable their movement. Without significant energy input into the system, the electrons remain immobile, and the material does not conduct electricity.

Therefore, we can conclude that the distinguishing characteristic of good insulators is the lack of freely moving electrons.

This question asks us to determine which object has the highest density. To do this, we’ll calculate the density for each object using the mass and volume values provided.

Density is given by the formula: Density = Mass / Volume.

For Object A, with a mass of 1.5 g and volume of 0.50 cm³:
DensityA = 1.5 g / 0.50 cm³
DensityA = 3.0 g/cm³.

For Object B with a mass of 3.0 g and volume of 0.75 cm³:
DensityB = 3.0 g / 0.75 cm³
DensityB = 4.0 g/cm³.

For Object C, with a mass of 4.5 g and volume of 1.00 cm³:
DensityC = 4.5 g / 1.00 cm³
DensityC = 4.5 g/cm³.

For Object D, with a mass of 6.0 g and volume of 1.50 cm³:
DensityD = 6.0 g / 1.50 cm³
DensityD = 4.0 g/cm³.

Its density is lower than that of Objects B, C, and D. This is because, despite its small volume, its mass is proportionally low, leading to a lower density.
Its density is higher than Object A but lower than Object C. Even though the mass of Object B increased more than its volume relative to Object A, it did not surpass the density of Object C.
From the above calculations:
Object A has a density of 3.0 g/cm³.
Object B has a density of 4.0 g/cm³.
Object C has a density of 4.5 g/cm³.
Object D has a density of 4.0 g/cm³.

Therefore, we can conclude that Object C has the highest density at 4.5 g/cm³.

Its density is the same as that of Object B but is less than Object C. Even though it has the highest mass among all objects, its larger volume prevents it from having the highest density.
This question asks us to identify the correct orbital geometry (also referred to as electron domain geometry) for an atom that is sp³ hybridized.

The concept of hybridization is grounded in the arrangement of electron domains around a central atom. An sp³-hybridized atom combines one s and three p orbitals from the central atom, leading to four equivalent hybrid orbitals. Orbitals are regions around an atom’s nucleus where electrons are likely to be found. Different types of orbitals, like s, p, d, and f orbitals, have distinct shapes and energies.
Relatedly, VSEPR theory predicts molecular shapes based on minimizing electron pair repulsions around a central atom. It considers the number of bonding and lone pairs of electrons around the central atom to predict the molecular geometry. The hybrid orbitals determined through hybridization play a role in this prediction by influencing the geometry of electron pairs around the central atom.

A linear geometry has only two electron domains, so it doesn’t correspond to sp³ hybridization.
A trigonal planar geometry is associated with three electron domains, which doesn’t match the four domains of sp³ hybridization.
When we consider examples like carbon in methane (CH4) or nitrogen in ammonia (NH3), both are sp3 hybridized. The four equivalent hybrid orbitals will orient themselves in space in such a way as to be as far apart from each other as possible. This results in them pointing toward the four corners of a tetrahedron.
Therefore, the correct orbital geometry for an sp3 hybridized atom is tetrahedral, making option C the right answer.
An octahedral geometry is based on six electron domains and does not fit the characteristics of sp³ hybridization.
This question asks us to identify which valence electron configurations indicate an atom in an excited state. In other words, we’re looking for an electron configuration that isn’t the typical ground state for an element.

Atoms have different energy levels, called orbitals, around their nucleus, where electrons reside. The Aufbau principle states that electrons first fill the lowest energy orbitals before moving to higher energy orbitals. In other words, electrons occupy orbitals in increasing order of their energy levels.

However, sometimes atoms absorb energy, like when something is heated. This extra energy makes electrons jump to higher energy levels temporarily. When electrons are in these higher levels, the atom is excited. But this state isn’t stable. The electrons eventually return to their original, lower energy levels, releasing the extra energy they absorbed as light.

An atom’s electron configuration outlines its electrons’ arrangement across various energy levels and orbitals. The combination of numbers and letters, such as 1s², 2s², 2p⁶, and so on, convey important information about this arrangement. The numbers signify the energy levels where electrons are situated. These energy levels progress from lower to higher energy levels. The accompanying letters, denoted as s, p, d, and f, indicate the distinct types of orbitals within each energy level where electrons can be found. For instance, the notation 1s² implies two electrons in the s orbital of the first energy level. Similarly, 2s² signifies the presence of two electrons within the s orbital of the second energy level.

This is the ground state electron configuration for beryllium (Be). The electrons are filled in the lowest energy orbitals available.
This is a ground-state electron configuration as the electrons are filled in the lowest energy orbitals available.
Therefore, we can conclude that option C (3s²3d¹) is correct, representing an atom in an excited state. This is because the configuration includes an electron in the 3d orbital before filling the 3p orbital, which contradicts the usual order of filling. This arrangement points to the atom absorbing extra energy, causing one electron to jump to a higher energy level than expected, resulting in an excited state.
This is consistent with the early transition metals in their ground state, where the 4s fills before the 3d.
The question aims to understand the trend in ionization energy as one moves from left to right across a period (horizontal row) on the periodic table. We must know that ionization energy is the energy required to remove an electron from a gaseous atom or ion.
There are a few important periodic table trends to be familiar with for the MCAT. As we move from left to right across a period in the periodic table:

The number of protons in the nucleus increases. This means there is an increasing positive charge in the nucleus, which tends to pull the electrons closer to the nucleus.

The atomic radius generally decreases. Even though electrons are being added, they are being added to the same principal energy level (same shell). The increase in protons pulls these added electrons closer to the nucleus, resulting in a smaller atomic radius.

The effective nuclear charge increases. Effective nuclear charge measures the net positive charge experienced by an electron in a multi-electron atom. As you move from left to right, there’s an increase in the number of protons (and a relatively small increase in electron-electron repulsion), which means each electron feels a stronger pull from the nucleus.

We can conclude that the ionization energy generally increases as the atomic number increases in a horizontal row of the periodic table, as it becomes more difficult to remove an electron due to the increased effective nuclear charge.
While it’s true ionization energy increases from left to right, it is not because of an increasing atomic radius. The atomic radius generally decreases as we move from left to right across a period.
Ionization energy does not decrease from left to right; it increases. Similarly, the effective nuclear charge does not decrease; it increases.
As mentioned before, ionization energy increases from left to right. Furthermore, the atomic radius generally decreases, not increases.
This question asks us to determine which pair of formulas represents caffeine’s empirical and molecular formulas.
The molecular formula indicates the actual number of atoms of each element present in a molecule of a compound. On the other hand, the empirical formula shows the simplest whole-number ratio of atoms of each element in the compound.

The first formula is overly simplified (it is not the lowest whole-number ratio), and the second doesn’t match the molecular formula of caffeine (it lacks two nitrogen atoms).
The first formula is correct as the empirical formula. However, the second formula should be the molecular formula, but it’s not different from the empirical formula here.
This option gives the molecular formula twice, offering no difference between the empirical and molecular formulas.
We can count the atoms in the provided image to get the molecular formula of caffeine as C₈H₁₀N₄O₂. This means that a caffeine molecule contains 8 carbon atoms, 10 hydrogen atoms, 4 nitrogen atoms, and 2 oxygen atoms.
We simplify the molecular formula to its lowest whole-number ratio to find the empirical formula. In this instance, each number in the molecular formula can be divided by 2:
For Carbon: C₈ ÷ 2 = C₄
For Hydrogen: H₁₀ ÷ 2 = H₅
For Nitrogen: N₄ ÷ 2 = N₂
For Oxygen: O₂ ÷ 2 = O (usually just written as O)
Therefore, we are left with the empirical formula: C₄H₅N₂O.
This question is focused on determining the element with the highest electron affinity among the given choices. We must know that electron affinity quantifies the energy change when an atom in its gaseous state gains an extra electron. At a fundamental level, electron affinity is influenced by the atomic structure, particularly the distance of the outermost shell from the nucleus and the number of electrons in this outermost shell.


Carbon is in Group 4 of the periodic table, meaning it has four valence electrons. This places it in a middle ground where it’s equally poised to gain or lose electrons to achieve a stable configuration. While it can gain electrons, its desire is less intense than fluorine’s. Plus, it’s earlier in the period (row), so it has less nuclear charge and less pull on additional electrons than elements further to the right.
Fluorine is part of Group 7 on the periodic table, known as the halogens. With seven valence electrons, it’s just one short of completing its outer electron shell.

The periodic trend for electron affinity increases as we move from left to right across a period, mainly due to increasing nuclear charge, which pulls the electrons closer, increasing attraction for additional electrons. Since fluorine is almost at the end of its period (row) and close to the noble gases (which have full outer shells and typically don’t want to gain or lose electrons), it has a strong inclination to gain that one additional electron to achieve a stable electron configuration similar to noble gases. This makes its electron affinity significantly high.

Oxygen resides in Group 6, holding six valence electrons. It wants to gain two more electrons to achieve a full outer shell. Although it’s closer to the noble gas configuration than carbon, it still doesn’t surpass fluorine in its desire to gain that extra electron. The increased nuclear charge from added protons as we move across the period does make oxygen have a higher electron affinity than carbon, but still not as high as fluorine.
Magnesium belongs to Group 2. These elements tend to lose their two valence electrons to achieve a noble gas configuration of the previous period. The energy change when magnesium gains an electron is not favorable. Instead, it’s more energetically favorable for magnesium to lose electrons. This nature makes its electron affinity low.
This question asks us to identify which metals can be paired with copper in a galvanic cell such that copper is reduced. We must know that a galvanic cell is a device that can produce electrical energy through a spontaneous redox (reduction-oxidation) reaction. The reaction’s spontaneity can be determined by the cell potential (Ecell°), which needs to be positive for the reaction to occur spontaneously.

Given the standard reduction potentials, we can compute the Ecell° for the galvanic cell pairing copper with each metal by using the formula:

Ecell° = Ecathode° – Eanode°

Where Ecathode° is the reduction potential of the substance getting reduced (in this case, copper) and Eanode° is the reduction potential of the substance getting oxidized.

Using the standard reduction potentials given: Ecell° = +0.52 V – (+0.80 V) = -0.28 V. Since this value is negative, silver cannot be paired with copper in a galvanic cell where copper is reduced. The negative cell potential means the reaction is nonspontaneous in the given direction. Thus, this choice is incorrect.
Although copper can form a galvanic cell with lead where copper is reduced (as shown by the positive Ecell° of +0.65 V), this answer choice is incorrect because it fails to mention zinc, which can also form a spontaneous galvanic cell with copper.
For the galvanic cell to be spontaneous, Ecell° must be positive. Now, let’s compute the Ecell° for each metal paired with copper:

1. Copper and Silver:
Ecell° = +0.52 V – (+0.80 V) = -0.28 V (Not spontaneous)

2. Copper and Lead:
Ecell° = +0.52 V – (-0.13 V) = +0.65 V (Spontaneous)

3. Copper and Zinc:
Ecell° = +0.52 V – (-0.76 V) = +1.28 V (Spontaneous)

Considering the above results, the metals that can be paired with copper in a galvanic cell such that copper gets reduced are lead and zinc, leading us to Answer Choice C: With lead and zinc.

While zinc is a correct pair for copper, as shown by the positive Ecell° of +1.28 V, silver cannot be paired with copper in a galvanic cell where copper is reduced, as explained in the evaluation for Answer Choice A. Therefore, even though part of this choice is accurate, the inclusion of silver makes it incorrect.
This question revolves around the concept of isoelectronic species. Isoelectronic species have the same number of electrons, although they can be different atoms or ions. To determine if something is isoelectronic with neon (Ne), we must understand how many electrons neon has and compare that count with the other given species.
Helium (He) is also a noble gas but from the first period of the periodic table. It has a full valence shell with 2 electrons. Thus, with only 2 electrons, helium is not isoelectronic with neon.

Fluorine (F) has 10 electrons. Magnesium (Mg2+) has 10 electrons after losing two. Sodium (Na+) has 10 electrons after losing one.

Therefore, helium is the only one not isoelectronic with neon.

Fluorine’s atomic structure is such that it’s one electron away from having a full valence shell. Gaining an electron and becoming the fluoride ion (F) achieves a more stable configuration that mirrors neon’s electron distribution.
Magnesium ion, after shedding two electrons, reaches a stable electron configuration resembling neon’s.
Sodium, by losing a single electron, accomplishes a stable electron configuration resembling neon’s.
This question seeks to determine which of the two gases diffuses faster based on their respective densities. The behavior of gases and their diffusion rates can be explained using Graham’s Law of Effusion. This law states that the rate of effusion (or diffusion) of a gas is inversely proportional to the square root of its molar mass. In simpler terms, lighter gases diffuse more quickly than heavier ones.
Since gas X has a lower density than gas Y, it also has a lower molar mass, and thus, diffuses faster than gas Y.
As per Graham’s Law, a gas with a higher molar mass would diffuse slower, not faster, so the reasoning behind the selection of gas X here is flawed.
This statement is incorrect because gas Y is denser and therefore has a higher molar mass than gas X, not a lower one.
While it is true that gas Y has a higher molar mass than gas X, this would mean gas Y diffuses more slowly, not faster. Therefore, this choice is also incorrect.
The question revolves around catalysis and its effect on chemical equilibrium. A catalyst increases the rate of reaction by lowering activation energy but does not alter the equilibrium position. It helps the system reach equilibrium faster but doesn’t change the concentrations of products or reactants at equilibrium.
Although a catalyst will accelerate the rate at which the reaction reaches equilibrium, it won’t change the amount of ammonia (or any other component) at equilibrium.
Introducing a catalyst will not change the amount of ammonia present at equilibrium.
This option is also incorrect for the same reason as Choice A. A catalyst doesn’t change the equilibrium position; it only helps the system achieve it faster.
The equilibrium constant (K) gives the ratio of products to reactants at equilibrium, but a catalyst doesn’t alter this ratio. Therefore, the equilibrium constant’s value remains unaffected by a catalyst’s presence.

The key principle underlying this question is the concept of solubility equilibrium and how it’s affected by the pH of the solution. We must recall that an ionic compound’s solubility can be influenced by the pH of the solution due to the presence of its basic or acidic ions.

When we discuss the solubility of an ionic substance in a solution, it refers to the maximum amount of that substance that can dissolve in the solution. The dissolution of an ionic substance leads to the release of its constituent ions. Some of these ions can interact with the solution, changing the equilibrium and, therefore, solubility.

AgCl doesn’t have a notably basic or acidic anion or cation. Thus, its solubility is not likely to be significantly impacted by either an acidic or basic solution.
When we have a basic anion, like OH in the case of Pb(OH)2, it can react with acidic protons (H+) in the solution. This reaction consumes the OH ions, pulling the equilibrium to the right and enhancing solubility following Le Chatelier’s principle.

This shift means more Pb(OH)2 will dissolve to replace the consumed OH ions. Because HCl is an acid and increases the concentration of H+ ions in the solution, Pb(OH)2 is more likely to dissolve in an HCl solution because of this reaction between H+ ions and OH ions.

The anion in CaF2 is F, which isn’t particularly basic. It’s not as likely as OH to react with H+ ions. Therefore, CaF2 solubility would not be significantly impacted by the pH of the solution.
HI is a strong acid. In an acidic or basic solution, it would readily dissociate into H+ and I ions. However, its solubility is not primarily influenced by the pH of the solution, unlike an ionic substance with a basic or acidic ion.

Electrolytic cells are a type of electrochemical cell designed to drive non-spontaneous chemical reactions forward using an external electrical source (such as a battery). These cells are crucial in industries for processes such as electroplating and the extraction or purification of metals.
The primary species being reduced is Cu2+, so increasing the sulfate concentration would have no positive effect on the deposition rate of copper.
In an electrolytic cell, the cathode is the site of reduction, where cations gain electrons, while the anode is the site of oxidation, where anions lose electrons. For the production of pure copper from CuSO4, the half-reaction at the cathode is:

Cu2+ (aq) + 2e ⇒ Cu(s)

The rate of this reduction (and, similarly, the corresponding oxidation at the anode) is intrinsically tied to the number of electrons flowing per unit time, essentially the definition of electric current. The larger the current, the greater the number of electrons available to reduce Cu2+ ions to metallic copper per unit time, thereby increasing the production rate.

In other words, current is the driving force in electrolytic processes. Ohm’s Law (I = V/R), which states that the current (I) in a circuit is proportional to the voltage (V) and inversely proportional to the resistance (R), is foundational here. In an electrolytic cell, the voltage supplied will overcome the cell potential and any resistance (due to the electrolyte and other components), ensuring electron flow. By increasing the current, we are maximizing the number of electrons available for the reduction reaction and, therefore, maximizing the rate of copper deposition.

While the size of the cathode dictates the available surface area for reduction reactions, without an increase in current, just increasing surface area will not inherently speed up the rate of copper production. A larger cathode would allow for more simultaneous reduction events if current and Cu2+ concentration are sufficiently high. Since this answer requires a few more steps of justification than option B, we can eliminate this answer choice.
The anode’s size similarly dictates the available surface area for oxidation reactions. However, in the context of this specific reaction, where we are focusing on the production of copper, the anode’s size is less important than the current in determining the rate of copper deposition (which takes place at the cathode).

This question asks us to determine where the 18O atom will be found in the product of the provided schematic. We must be able to track where the oxygen moves as the reaction progresses.
Position a remains unchanged in the reaction and does not incorporate the 18O label from the methanol.
In the reaction mentioned, the hemiacetal is treated with 18O-labeled methanol. We can break down each step of the reaction:
1. The hydroxyl group (from the hemiacetal) is protonated by the acid.
2. This creates a water molecule, which is a good leaving group. It departs, forming a carbocation.
3. The 18O-labeled methanol attacks this carbocation, leading to the formation of the acetal. The oxygen from the labeled methanol becomes directly attached to the cyclic carbon bonded to the ring’s oxygen.

Therefore, the 18O label from methanol will appear where the methoxy group is attached, which is position b. Position a comes from the initial hemiacetal and remains unchanged. Position c is in water and will not contain the 18O label since it originates from the hemiacetal OH group.

This would imply that the 18O label from the methanol ends up in the water molecule. This is inaccurate, as the water molecule originates from the hydroxyl group of the hemiacetal.
This implies that both the methoxy group and the water molecule have the 18O label, but as explained, only the methoxy group will have this label.

From the question stem and a glance at the answer options, we can see that this question asks us to determine the classification of compound X. We must understand peptide formation and nomenclature to answer this question.

Peptides are essentially short chains of amino acids, which are the building blocks of proteins. The bond that connects two amino acids is called a peptide bond, formed through a dehydration synthesis reaction between the carboxyl group of one amino acid and the amino group of another. This results in the elimination of a molecule of water.

This would be accurate if Compound X had only two amino acids connected by a single peptide bond. However, that is not the case.
Logically, a dipeptide cannot have two peptide bonds since it only contains two amino acids.
To determine the classification of Compound X, we must look to see how many amino acids it contains. Monopeptides contain only one amino acid and, therefore, have no peptide bonds. Dipeptides are made of two amino acids connected by one peptide bond. Tripeptides are comprised of three amino acids interconnected by two peptide bonds. Tetrapeptides are composed of four amino acids and have three peptide bonds.

The figure in the question shows that Compound X consists of three amino acid residues (phenylalanine, valine, and alanine). With this information, we can deduce that two peptide bonds are holding these three residues together (one between phenylalanine and valine and the other between valine and alanine).

This suggests Compound X is made up of four amino acids, which is inconsistent with the figure provided. Instead, we see three amino acids phenylalanine, valine, and alanine (from left to right).

This question asks us to determine what functional groups are found in benzoin. Fortunately, we are given the molecular formula of benzoin, so we must simply identify the functional groups present.

Functional groups are atomic configurations within molecules responsible for specific chemical behaviors. Recognizing them is crucial for predicting the reactivity and properties of organic compounds and is therefore heavily tested on the MCAT.

Carboxylic acids are a type of organic compound characterized by a carboxyl group (-COOH) where a carbonyl group (C=O) and a hydroxyl group (-OH) are both bonded to the same carbon atom. This arrangement imparts acidic properties to these compounds. However, this particular structural feature is absent in benzoin. Unlike carboxylic acids, the carbonyl and hydroxyl groups are not directly bonded to the same carbon atom in benzoin.
Ethers are organic compounds characterized by an oxygen atom (-O-) bonded to two carbon atoms. This oxygen atom forms a bridge between the carbon atoms, resulting in an oxygen-carbon-oxygen linkage. Benzoin, however, does not exhibit this specific arrangement. Instead, benzoin consists of two aromatic rings connected by a carbonyl group (C=O), and the oxygen atom does not bond to two carbons, as is the case in ethers.
An aldehyde is an organic compound with a carbonyl group (C=O) bonded to a carbon atom and a hydrogen atom. Benzoin does not have the structure of an aldehyde. Rather, it consists of two aromatic rings linked by a carbonyl group (C=O) without the direct bonding of a hydrogen atom to the carbon of the carbonyl group, which is characteristic of aldehydes.
In benzoin, we can identify two primary functional groups: (1) alcohol denoted by the -OH group. Alcohols are characterized by a hydroxyl group attached to a saturated carbon atom. This group plays a vital role in the reactivity of alcohols and provides specific physical properties like increased boiling points due to hydrogen bonding. (2) A ketone characterized by a carbonyl group C=O flanked by two carbons. Their unique carbonyl configuration grants ketones distinct reactivity patterns, differentiating them from other carbonyl-containing compounds.

From the given molecular structure of benzoin and the available answer choices, it is clear that the molecule only possesses a ketone functional group.