57
doi.org/10.26434/chemrxiv.8862677.v1 Catalytic Resonance Theory: SuperVolcanoes, Catalytic Molecular Pumps, and Oscillatory Steady State M. Alexander Ardagh, Turan Birol, Qi Zhang, Omar Abdelrahman, Paul Dauenhauer Submitted date: 12/07/2019 Posted date: 15/07/2019 Licence: CC BY 4.0 Citation information: Ardagh, M. Alexander; Birol, Turan; Zhang, Qi; Abdelrahman, Omar; Dauenhauer, Paul (2019): Catalytic Resonance Theory: SuperVolcanoes, Catalytic Molecular Pumps, and Oscillatory Steady State. ChemRxiv. Preprint. Catalytic reactions on surfaces with forced oscillations in physical or electronic properties undergo controlled acceleration consistent with the selected parameters of frequency, amplitude, and external stimulus waveform. In this work, the general reaction of reversible A-to-B chemistry is simulated by varying the catalytic (heat of reaction, transition state and intermediate energies) and oscillation parameters (frequency, amplitude, endpoints, and waveform) to evaluate the influence on the overall catalytic turnover frequency and steady state extent of conversion. Variations of catalytic cycle energies are shown to comprise a superVolcano of superimposed individual Balandin-Sabatier volcano plots, with variations in linear scaling relationships leading to unique turnover frequency response to forced oscillation of the catalyst surface. Optimization of catalytic conditions identified a band of forced oscillation frequencies leading to resonance and rate enhancement as high as 10,000x above the static Sabatier maximum. Dynamic catalytic reactions conducted at long times achieved oscillatory steady state differing from equilibrium consistent with the imposed surface oscillation amplitude acting as a ‘catalytic pump’ relative to the Gibbs free energy of reaction. File list (2) download file view on ChemRxiv Dynamics_2_Manuscript_ChemRxiv_ver_04.pdf (2.08 MiB) download file view on ChemRxiv Supplementary_Information_ver_03.pdf (545.56 KiB)

Catalytic Resonance Theory: SuperVolcanoes, Catalytic

  • Upload
    others

  • View
    22

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

doi.org/10.26434/chemrxiv.8862677.v1

Catalytic Resonance Theory: SuperVolcanoes, Catalytic MolecularPumps, and Oscillatory Steady StateM. Alexander Ardagh, Turan Birol, Qi Zhang, Omar Abdelrahman, Paul Dauenhauer

Submitted date: 12/07/2019 • Posted date: 15/07/2019Licence: CC BY 4.0Citation information: Ardagh, M. Alexander; Birol, Turan; Zhang, Qi; Abdelrahman, Omar; Dauenhauer, Paul(2019): Catalytic Resonance Theory: SuperVolcanoes, Catalytic Molecular Pumps, and Oscillatory SteadyState. ChemRxiv. Preprint.

Catalytic reactions on surfaces with forced oscillations in physical or electronic properties undergo controlledacceleration consistent with the selected parameters of frequency, amplitude, and external stimuluswaveform. In this work, the general reaction of reversible A-to-B chemistry is simulated by varying the catalytic(heat of reaction, transition state and intermediate energies) and oscillation parameters (frequency, amplitude,endpoints, and waveform) to evaluate the influence on the overall catalytic turnover frequency and steadystate extent of conversion. Variations of catalytic cycle energies are shown to comprise a superVolcano ofsuperimposed individual Balandin-Sabatier volcano plots, with variations in linear scaling relationshipsleading to unique turnover frequency response to forced oscillation of the catalyst surface. Optimization ofcatalytic conditions identified a band of forced oscillation frequencies leading to resonance and rateenhancement as high as 10,000x above the static Sabatier maximum. Dynamic catalytic reactions conductedat long times achieved oscillatory steady state differing from equilibrium consistent with the imposed surfaceoscillation amplitude acting as a ‘catalytic pump’ relative to the Gibbs free energy of reaction.

File list (2)

download fileview on ChemRxivDynamics_2_Manuscript_ChemRxiv_ver_04.pdf (2.08 MiB)

download fileview on ChemRxivSupplementary_Information_ver_03.pdf (545.56 KiB)

Page 2: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 1

Catalytic Resonance Theory: SuperVolcanoes, Catalytic Molecular Pumps,

and Oscillatory Steady State

M. Alexander Ardagh1,2, Turan Birol1, Qi Zhang1, Omar Abdelrahman2,3, Paul J. Dauenhauer1,2*

1University of Minnesota, Department of Chemical Engineering and Materials Science, 421 Washington

Ave SE, Minneapolis, MN 55455 2Catalysis Center for Energy Innovation, University of Delaware, 150 Academy Street, Newark, DE

19716 3University of Massachusetts Amherst, Department of Chemical Engineering, 159 Goessmann

Laboratory, 686 North Pleasant Street, Amherst, MA 01003.

*Corresponding author: [email protected]

1.0 Introduction. The efficient catalytic

transformation of feedstocks to chemicals and

materials for society remains at the forefront of

technological needs in the 21st century.[1] Grand

challenges in catalysis aim for the sustainable

manufacture of monomers[2,3,4], the elimination of

carbon dioxide[5,6], the fixing of nitrogen[7,8,9,10], and

the environmental remediation of pollutants[11,12].

New catalysts should be synthesized from low-cost,

earth-abundant materials[13] that operate with long-

term stability and negligible environmental impact

within a process designed to optimally benefit from

catalytic rate enhancement. Catalyst development

also aims for maximum catalytic rate within the

limitations of the Sabatier principle[14] for multi-

step chemistry which includes adsorption, surface

catalytic reactions, and desorption (Figure 1A).

Catalyst design has historically focused on

selecting the optimum catalytic active site within a

material which provides a balance of strong and

weak adsorbate binding energies[15]. Strong

binding is rate limiting in desorption resulting in

rate inhibition from the product, while weak

binding is limiting in dissociative adsorption or

surface reaction[16,17,18]. The original strategy in

these systems tuned catalyst characteristics to exist

at the maximum possible turnover frequency (i.e.,

the Sabatier volcano peak)[19]. More recent

strategies aim to shift the volcano peak via

alteration of the linear scaling relations, which

describe the predicted surface activation energy as

a function of surface reaction enthalpy[20,21,22,23].

Varying of the linear scaling relations of surface

reaction activation energies has the potential to

Abstract. Catalytic reactions on surfaces with forced oscillations in physical or electronic properties

undergo controlled acceleration consistent with the selected parameters of frequency, amplitude, and

external stimulus waveform. In this work, the general reaction of reversible A-to-B chemistry is

simulated by varying the catalytic (heat of reaction, transition state and intermediate energies) and

oscillation parameters (frequency, amplitude, endpoints, and waveform) to evaluate the influence on the

overall catalytic turnover frequency and steady state extent of conversion. Variations of catalytic cycle

energies are shown to comprise a superVolcano of superimposed individual Balandin-Sabatier volcano

plots, with variations in linear scaling relationships leading to unique turnover frequency response to

forced oscillation of the catalyst surface. Optimization of catalytic conditions identified a band of forced

oscillation frequencies leading to resonance and rate enhancement as high as 10,000x above the static

Sabatier maximum. Dynamic catalytic reactions conducted at long times achieved oscillatory steady

state differing from equilibrium consistent with the imposed surface oscillation amplitude acting as a

‘catalytic pump’ relative to the Gibbs free energy of reaction.

Page 3: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 2

change the shape of Sabatier volcano plots and shift

the maximum possible turnover frequency.

We have recently described an alternative

concept for catalyst enhancement in the form of

surface adsorbate binding oscillation[24]. By this

approach depicted in Figure 1B, the catalyst surface

undergoes periodic external stimulation (e.g.

electric field, strain) such that reaction surface

intermediates experience oscillating binding energy

with time. The amplitude of the oscillating

stimulus generates variation in surface species

binding energy, with the endpoints of the oscillation

amplitude dictated by the combination of the

selected catalyst surface and the condition of the

external stimulus. Stimulus waveforms can also

vary in frequency, type (e.g. square, sinusoidal),

and periodicity (e.g. combinations of waveforms).

For the reversible A to B reaction, the highest

catalytic turnover frequency occurs over a band of

imposed frequencies which resonate with the

natural frequencies of the catalytic surface reaction.

The turnover frequency response of the

catalytic system to external oscillating stimuli

depends on the kinetics of the surface mechanism

and the resulting shape of the Balandin-Sabatier

volcano curve[25,26]. A catalytic system may exhibit

a linear Brønsted-Evans Polanyi (BEP) relationship

with slope, α, and intercept, β, relating the surface

reaction thermodynamics to the surface reaction

transition state. However, another consideration in

volcano plot architecture is the dynamic response

of individual surface species (e.g., A*, B*) to

external stimulus. For variation of any descriptor

(e.g. binding energy of an intermediate or single

atom), surface species exhibit differing binding

energies with the surface. In our previous work on

surface resonance of the reversible A to B

reaction[24], the considered catalytic system

exhibited two-fold variation in the binding energy

of B* relative to A* for the same extent of external

stimulus. A broader understanding of this ratio

(referred to here as gamma, γ) on the structure of

volcano plots is required to predict the potential of

imposed external dynamic stimulus across a broad

range of catalytic chemistries and materials.

In addition to rate, chemical equilibrium

between the reactants and products limits many

catalytic systems. As shown in Figure 1C,

industrially important reactions including ammonia

synthesis, water-gas shift chemistry, dry reforming

of methane, methanol synthesis, and alkane

dehydrogenation all exhibit Gibbs free energy of

reaction (and associated equilibrium constants, Ki)

at standard conditions limiting high overall

conversion[27,28,29,30,31]. To overcome these inherent

limitations, equilibrium between reactants and

products can be manipulated via reaction

conditions; ammonia synthesis is operated at high

pressure[32,33], while water-gas-shift chemistry is

conducted in staged reactors of varying

temperature[34,35]. Other strategies for overcoming

equilibrium limitations combine a second reaction

or separation phenomenon to drive a reaction. For

example, alkane dehydrogenation is combined with

oxidation[36], ammonia synthesis is combined with

absorption[37,38], and carbon-carbon coupling

chemistries such as cycloaddition can be combined

with dehydration[39,40].

1

10

100

50 160 270 380 490 600

Equili

brium

Convers

ion [

%]

Temperature [ C]

100 atm

N2 + 3H2 ↔ 2NH3

CO2 + H2 ↔ CO + H2O

1.0 atm

1.0 atm: CO2 + CH2 ↔ 2CO + 2H2

A

A*

B

B*

ΔHA ΔHB

EA

CA

-5.00

-4.00

-3.00

-2.00

-1.00

0.00

1.00

2.00

3.00

4.00

5.00

-0.50 -0.25 0.00 0.25 0.50 0.75 1.00

Tu

rno

ver

Fre

qu

en

cy

to

B,

TO

FB

[s-1

]

Relative Binding Energy of B [eV]

Amplitude

Accessib

le

Reaction

Rate

s

Desorption Rate

Surface Reaction

Rate

XA~1%105

104

103

102

101

1

10-1

10-2

10-3

10-4

10-5

B

Figure 1. Kinetic and Thermodynamic Challenges of Catalytic Reactions. (A) Gas-phase chemical species A

and B react on catalyst surfaces through adsorbed surface species A* and B* with forward activation energy, Ea. (B)

Conversion of A-to-B general chemicals volcano curve operating at 1% conversion with oscillating binding energy

of B* with resonance rate in purple. (C) Reaction equilibrium of ammonia synthesis (red), reverse water-gas-shift

(blue), and dry reforming of methane (green).

Page 4: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 3

The driving force for chemical reaction is the

difference in Gibbs free energy between reactant

and product, with a reaction system stabilizing at

equilibrium defined as the two states existing at

equal Gibbs free energy. The two strategies for

manipulating reactions are based on this definition;

change the conditions and equilibrium, or deprive

the system of one of the two components (reactant

or product) to unbalance the free energy

distribution (i.e., Le Chatelier’s principle)[41,42]. A

third strategy for driving physical systems away

from equilibrium is the application of work; added

energy can perturb a physical system to a steady-

state condition different from equilibrium such as

the case of electrocatalytic water splitting[43,44] or

electrocatalytic ammonia synthesis[45].

Alternatively, a catalyst surface with dynamically-

oscillating adsorbate binding energy provides work

to a catalytic reaction to raise the energy of

adsorbed surface species[46]. The imposed

amplitude of surface binding energy oscillation (0.1

< ΔU < 1.5 eV) serves as a ‘catalytic pump’ to raise

a surface adsorbate to a higher energy state. This

surface energy input manifests as a rate

enhancement and a deviation from equilibrium of

the surface reaction, with the extent of variation in

reaction rate and conversion depending on the type

of catalytic chemistry as well as the parameters

associated with the imposed oscillating surface

stimulus.

Here we will map out the different structures of

Sabatier volcano plots based on definitions of the

five key parameters of single-reaction (A-to-B)

reversible catalytic systems. Example systems

representative of the different kinetic regimes

associated with different kinetic parameter

combinations are simulated as catalytic reactors to

explore the ability of the dynamically oscillating

catalyst surface to preferentially promote catalytic

turnover frequency. Selected systems are evaluated

within the context of the forced oscillator

parameters of imposed surface binding energy

frequency, amplitude, amplitude position, and

oscillating waveform. Specific conditions are

identified leading to surface resonance and

enhanced overall catalytic rate relative to the

Sabatier maximum, while other conditions

preferentially promote a steady state condition

differing from equilibrium. The relationship

between the applied oscillation energy (i.e., surface

work) and the resulting oscillatory steady state is

then evaluated to understand the conditions leading

to tunable reaction conversion with broad

application.

2.0 Computational Methods. Computational

simulations were conducted using Matlab 2017b

and 2019a, as well as the supercomputing resources

at the Minnesota Supercomputing Institute (MSI).

Reactor and dynamic catalysis codes are provided

in SI section S1. A model reversible reaction (A ↔

B) was implemented as a gas-phase catalytic

reaction system with three reversible elementary

steps: (i) reversible adsorption of A, (ii) surface

forward and reverse reaction of A* ↔ B*, and (iii)

reversible desorption of B. Pre-exponential factors

were set to constant values typical for each type of

elementary step. For adsorption steps, a pre-

exponential of 106 (bar-s)-1 was selected, while a

pre-exponential of 1013 s-1 was used for surface

reaction and desorption steps. Surface reaction

activation energies (Ea,sr) were calculated based on

the specified Brønsted-Evans-Polanyi (BEP)

relationship parameters and the heat of reaction for

the surface reaction (ΔHsr)[47], as shown in Equation

1. The activation energy of adsorption was set to 0

kJ mol-1, and the activation energy of desorption

was set to the binding energy for each species,

𝐸𝑎,𝑠𝑟 = 𝛼𝛥𝐻𝑠𝑟 + 𝛽 (1)

where α is the BEP proportionality constant, and β

is the BEP offset.

The reaction chemistry was specified by the

overall gas-phase heat of reaction (ΔHovr), species

binding energies (BEA and BEB), and the BEP

relationship parameters (α and β). Vapor-phase

flow reactors (CSTR) and batch reactors were

modeled by systems of ordinary differential

equations (ODEs) with two gas phase species, A

and B, and two surface species, A*, B*, with open

site *. Differential equations for the CSTR were as

follows:

𝑑[𝐴]

𝑑𝑡=�̇�

𝑉([𝐴]𝑓 − [𝐴]) − 𝑘1,𝑓[𝐴]𝑅𝑇𝜃

∗ 𝑁𝑠𝑖𝑡𝑒𝑠

𝑉+

𝑘1,𝑟𝜃𝐴∗ 𝑁𝑠𝑖𝑡𝑒𝑠

𝑉 (2)

𝑑[𝐵]

𝑑𝑡=�̇�

𝑉([𝐵]𝑓 − [𝐵]) − 𝑘3,𝑟[𝐵]𝑅𝑇𝜃

∗ 𝑁𝑠𝑖𝑡𝑒𝑠

𝑉+

𝑘3,𝑓𝜃𝐵∗ 𝑁𝑠𝑖𝑡𝑒𝑠

𝑉 (3)

Page 5: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 4

𝑑𝜃𝐴∗

𝑑𝑡= 𝑘1,𝑓[𝐴]𝑅𝑇𝜃

∗ − 𝑘1,𝑟𝜃𝐴∗ − 𝑘2,𝑓𝜃𝐴

∗ + 𝑘2,𝑟𝜃𝐵∗

(4)

𝑑𝜃𝐵∗

𝑑𝑡= 𝑘3,𝑟[𝐵]𝑅𝑇𝜃

∗ − 𝑘3,𝑓𝜃𝐵∗ + 𝑘2,𝑓𝜃𝐴

∗ − 𝑘2,𝑟𝜃𝐵∗

(5)

where [A] and [B] are gas phase concentrations in

M, �̇� is the volumetric flow rate in liters/s, V is the

reactor bed volume in liters, k’s are rate constants

in (bar-s)-1 or s-1, θ’s are surface coverages, and

Nsites is the number of catalytic active sites. In

addition, surface coverage is constrained by the site

balance shown in Equation 6.

1.0 = 𝜃∗ + 𝜃𝐴∗ + 𝜃𝐵

∗ (6)

For batch reactors, the flowrate terms from

Equations 2 and 3 are removed.

Differential equations were solved using built-

in Matlab ODE solvers ODE15s and ODE23tb

along with the “Radau” 3rd party Matlab ODE

solver. Tight solver tolerances of 10-6-10-8 relative

tolerance and 10-9 absolute tolerance were used due

to the ultra-stiff nature of the static and dynamic

catalysis systems. Performance comparison of the

various Matlab ODE solvers can be found in SI

section S2. Computational times were measured

using the ‘tic’ and ‘toc’ commands which act as a

stopwatch timer. The stopwatch was initiated right

before dynamic catalysis was applied in a

simulation and stopped after the last dynamic

catalysis iteration.

Dynamic catalysis was implemented by

varying the binding energies of A and B using

square, sinusoidal, triangle, and sawtooth shaped

waveforms. Waveforms were defined by their

oscillation amplitude ΔU in electron volts (eV),

oscillation frequency f in hertz (Hz), and oscillation

endpoints in eV. Oscillation amplitude is noted

relative to either surface intermediate as ΔUA or

ΔUB. The binding energies of A and B were varied

using a defined proportionality constant, γ,

according to Equation 7,

𝛥𝐵𝐸𝐵 = 𝛾𝛥𝐵𝐸𝐴 (7)

where the change in the binding energy of B was

specified using one of the four different waveform

equations (e.g. sinusoidal). A second parameter, δ,

was defined for the relationship between binding

energies. At any binding energy of B, the binding

energy of A was calculated using Equation 8:

𝐵𝐸𝐴 = (𝐵𝐸𝐵 − (1 − 𝛾)𝛿 − 𝛥𝐻𝑜𝑣𝑟)/𝛾 (8)

Heat maps of catalytic turnover frequency to B

(TOFB) were created in Matlab 2019a by

interpolating over discrete data sets obtained at

varying oscillation amplitudes, oscillation

frequencies, and oscillation endpoints. The

modified Akima cubic Hermite or spline

interpolation method was used to fit data sets and

the ‘smoothdata’ function was employed with the

moving average or Savitzky-Golay filter method to

average out overshoots due to interpolation.

Complete tabulated data sets of all heat maps are

provided in the supporting information.

3.0 Results and Discussion. As shown in Figure

2A, a general reaction enthalpy diagram for

chemical species A reversibly converting to B

forms intermediates A* and B* with transition

ΔHrxn

ΔHBEa = f(α,β)

ΔHsurf = f(γ,δ)

δ

A A* TS B* B

State 3

State 2

State 1

Heat of Adsorption of A, ΔHA

He

at o

f Adso

rptio

n o

f B, Δ

HB

(δ - ΔHrxn)

δSlope, γ

= 𝐻𝐵 𝐻𝐴

1.0

Figure 2. Parameters of Dynamic Heterogeneous Catalysis. Left. State-energy diagram of oscillating

heterogeneous catalyst. Right. Variation of the binding energy of B* linearly scales with the binding energy of A*

with slope, γB-A, and common point, δ.

Page 6: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 5

state, TS, in between. The overall heat of reaction,

ΔHovr, remains fixed for all conditions of catalyst

surface dynamics. Two more variables are required

to define the transition state (TS) energy, Ea, which

will vary with the surface intermediate energies. As

stated in Equation 1, the existing BEP relationships

with slope, α, and intercept, β, fully define all

possible transition state energies for catalytic

systems exhibiting this linear relationship.

Two new variables are required to define the

variation of surface intermediate energies. Due to

the differences in electronic and steric interactions

with the surface, variation of surface adsorbate

binding energies with external stimuli will not

affect all surface adsorbates equally. As depicted in

Figure 2A, the binding energy of A* changes less

than B* between each of the three potential states

resulting from external stimulus. To describe this

change differential, the extent of change is defined

to be gamma equal to the difference in binding

energy of B* divided by the difference in binding

energy of A*,

𝛾𝐵−𝐴 = 𝛥𝛥𝐻𝐵

𝛥𝛥𝐻𝐴 (9)

Additionally, there exists one condition (purple in

Figure 2A) leading to A* and B* existing at the

same surface energy; that energy is identified in

Figure 2A as delta, δ [eV].

If the external stimulus of a catalytic surface

imposes a linear response in the binding energy of

the surface adsorbates, then potential dynamic

catalysis scenarios can be depicted in the right panel

of Figure 2B. As shown, the binding enthalpies of

A* and B* are related by a line with slope, gamma

γB-A. It is important to note that gamma is defined,

and consideration of the reaction in reverse will

produce a gamma of inverse value. The point of

common surface energy, delta δ, exists for the

enthalpy of adsorption of B offset by the heat of

reaction. More complicated relationships (i.e., non-

linear) between the surface energies of intermediate

species require more than two variables, depending

on the applied model.

3.1 External Catalytic Stimuli. The ability to

tune adsorption enthalpy (also referred to as

binding energy, BE = -ΔHads), requires external

stimulus to modify the catalyst, adsorbate, or

interacting surface bond. One method of catalyst

stimulus is surface strain, which has been shown in

static systems (e.g., films on lattice-mismatched

substrates) to shift the metal d-band center and alter

the adsorption enthalpy of surface

adsorbates[48,49,50,51]. Electric fields applied to

adsorbed species can also tune the binding energy

of adsorbates with benefits for controlling and

accelerating catalytic surface chemistry, including

externally imposed fields or internal fields arising

in electrocatalytic reactors[52,53,54]. Other strategies

aim to alter the electronic structure of surface metal

atoms, such as field-effect modulation with a

dielectric and imposed back gate-voltage; this

method moves charge carriers into the surface

catalyst layer (e.g. metal or metal oxide) to shift the

d-band center and ultimately the interaction with

surface adsorbates[55,56, 57,58]. The impact of these

techniques on the relative changes in surface

intermediate energies and variation in the gamma-

delta plot (e.g., Figure 2B) of two surface

adsorbates depends with each combination of

surface intermediates, surface material/facet, and

type of imposed stimulus which so far has been

primarily explored via computation.

An initial comparison across surface

intermediate binding energies can be made by

comparing the energies between metals and metal

surface planes. For the case of ammonia, the

surface adsorption enthalpy of N*, NH*, NH2*, and

NH3* have been calculated by Mavrikakis and co-

workers on Re(0001)[59], Ru(0001)[60], Pt(111)[61],

Pd(111)[62], and Au(111)[63]. As shown in Figure

3A and 3B, the adsorption enthalpies of the various

intermediates are plotted relative to one another

with the gamma values of the intermediates

apparent in the fitted linear slopes. In the

comparison of NH* and NH2*, the binding energy

of NH* changes almost twice as much as the

binding energy of NH2* (slope ~ γNH2-NH ~ 0.53).

Similarly, in the comparison of NH* and NH3*, the

binding energy of NH* changes almost five times

as much as the binding energy of NH3* (γNH3-NH ~

0.21). These comparisons exist across different

metals, which is not a realistic scenario for

implementing dynamic catalysis (i.e., metals

cannot be periodically interchanged). Only

external stimulus on a single catalyst surface is

likely to physically achieve the conditions

necessary for dynamic catalytic resonance and rate

promotion.

Comparison of the potential for tuning

adsorbate binding energy via external stimuli are

presented for imposed electric field (Figure 3C) and

Page 7: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 6

surface strain (Figure 3D). As calculated by

McEwen and co-workers[64], electric fields ranging

from -1.0 to +1.0 V/Å applied to a Nickel surface

resulted in significant variation of the binding

energy of methane (blue in Figure 3C), hydrogen

(black), and formaldehyde (red) relative to

methanol. From the slopes of a linear best fit,

methane varied less with changes in methanol

(slope ~ γCH4-CH3OH ~ 0.18) than hydrogen (γH2-CH3OH

~ 0.51), while formaldehyde actually decreased in

binding energy as methanol increased in binding

energy (γCH2O-CH3OH ~ -1.31). Another method of

varying surface binding energy of surface strain is

presented in Figure 3D using the calculated binding

energies of Shengchun Yang and co-workers[65]. As

depicted in Figure 3D, the binding energy of carbon

monoxide and atomic oxygen were calculated for

the lowest energy adsorption geometry on the

M(111) surface of five different metals: Cu (black),

Rh (green), Ir (purple), Pd (orange), and Pt (blue).

Binding energies varied due to imposed surface

strain of -3.0% to +3.0%; while this large surface

strain is likely not physically feasible, it does

provide insight into the relative impact of strain on

oxygen relative to carbon monoxide. From these

calculations, there exists some initial evidence that

a two-parameter linear model can effectively

describe the relative change in binding energy

between two surface adsorbates.

3.2 Catalytic Dynamics - Definition. With the

definition of parameters in Figure 2 and the

calculations presented in Figure 3, it is possible to

completely define the surface catalytic kinetics for

a dynamic chemistry with five parameters: ΔHovr,

y = 0.53x - 0.12

-6

-5

-4

-3

-2

-1

0

-6 -5 -4 -3 -2 -1 0

He

at

of

Ad

so

rpti

on

, N

H2*

[eV

]

Heat of Adsorption, NH* [eV]

-1-2-3-4-5

-1

-2

-3

-4

-5Au(111)

Pd(111)

Pt(111)

Ru(0001)

Re(0001)

y = 0.21x + 0.34

-6

-5

-4

-3

-2

-1

0

-6 -5 -4 -3 -2 -1 0

Heat of Adsorption, NH* [eV]

He

at

of

Ad

so

rpti

on

, N

H3*

[eV

]

-1-2-3-4-5

-1

-2

-3

-4

-5

NH2* vs. NH* NH3* vs. NH*

A B

Au(111)

Pd(111)

Pt(111)

Ru(0001)

Re(0001)

-1.2

-1.0

-0.8

-0.6

-0.4

-0.2

0.0-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0

Ad

so

rpti

on

En

erg

y:

H2,

CH

2O

, C

H4

[eV

]

Adsorption Energy Methanol [eV]

CH2O

H2

CH4

E: -2.0 < V/Å < 2.0

C

-2.5

-2.0

-1.5

-1.0

-0.5-2.5 -2.0 -1.5 -1.0 -0.5

Ca

rbo

n M

on

ox

ide

Bin

din

g E

ne

rgy [

eV

]

Oxygen Binding Energy [eV]D

Pt

Cu

Rh

Ir

Pd

Figure 3. Gamma Parameters of Ammonia Synthesis. (A) The heat of adsorption of NH* and NH2* vary by γ ~

0.53. (B) The heat of adsorption of NH* and NH3* vary by γ ~ 0.21. Data for panels A and B calculated by Mavrikakis

and co-workers[59-63]. (C) The heat of adsorption of methanol varies differently than methane (CH4), hydrogen (H2),

and formaldehyde (CH2O) on Ni metal in the presence of an electric field varying between -1.0 to +1.0 V/Å calculated

by McEwen and co-workers[64]. (D) The binding energy of oxygen, O*, and carbon monoxide, CO*, on Copper

(black), Rhodium (green), Iridium (purple), Palladium (orange), and Platinum (blue) with strain of -3.0% to +3.0%[65].

Page 8: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 7

α and β to define the activation energy linear scaling

relationship, and γ and δ to define the linear

adsorption energy scaling relationship. Four

additional parameters exist related to the applied

dynamics of the catalyst binding energy oscillation

including: (1) surface oscillation frequency, f, (2)

surface oscillation amplitude, ΔU, (3) surface

oscillation endpoint, UE, and (4) surface oscillation

waveform shape (e.g., square, sinusoidal, etc.).

Additional parameters are required if the applied

surface stimulus is a composite of imposed

oscillations each with their own frequency and

amplitude. Finally, simulation of the dynamic

catalyst system requires all of the parameters

associated with the reactor including temperature,

pressure, reactant composition, and space time.

The introduction of dynamics for the simple

system of A reversibly reacting to B requires nine

variables for definition, which is more than double

the four parameters required to define the static

catalytic reaction (two adsorption enthalpies, one

activation energy, and overall reaction enthalpy).

As depicted in Figure 4A, as the complexity of

catalytic systems expands by the number of

catalytic surface reactions, the number of

parameters required to define the oscillating

systems significantly expands. Dynamic systems

require at least 2n + 1 parameters to define all

energetic states, where n is the number of

parameters required to define the static catalytic

reaction. If the scaling relationships to predict

changes in the transition state energy or the relative

binding energies of surface intermediates become

nonlinear, then the number of parameters further

expands accordingly.

The time required to achieve steady state in a

simulated continuous-flow stirred-tank reactor with

a dynamically operating catalyst increases orders of

magnitude from the simulation with a static

catalyst. As depicted in Figure 4B, the considered

A-to-B reversible catalytic reaction was simulated

using Matlab to achieve steady-state reaction

conditions in less than one second. However,

introduction of dynamics immediately increased

computational time to greater than a second for all

conditions. In particular, computational time

increased logarithmically above about 100 Hz such

that simulations above megahertz-imposed catalyst

binding energy waveforms required more than 104

seconds (~ 3.0 hrs).

Increased computational time is associated

with the large number of oscillations required to

achieve steady-state surface and reactor conditions.

As depicted in Figure 5A, a catalytic batch reactor

0

5

10

15

20

25

30

1 2 3 4 5

Num

ber

of

Reaction

Para

mete

rs [-]

Number of Surface Reactions [-]

Dynamic

Static

1.E-03

1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

1.E+05

1.E+06

1.E+07

1.E+08

1.E+09

1.E+10

1.E+11

1.E+12

1E-06 1E-03 1E+00 1E+03 1E+06 1E+09 1E+12

Co

mp

ute

r T

ime

[s]

Oscillation Frequency [Hz]

10-6 10-3 1 103 106 109 1012

10-2

1

10+2

10+4

10+6

10+8

10+10

10+12

Static

Dynamic, γ < 1

Dynamic, γ > 1

A B

Figure 4. Reaction Parameters of Static and Dynamic Surface Mechanisms. (A) The number of parameters

required to define all energies of all states in a catalytic surface reaction mechanism differ between a static, steady-

state reaction and a dynamic catalytic reaction with oscillating states. Reactor parameters (e.g., temperature, pressure,

composition, space time) are not included in this figure. (B) Computational time for static (black), dynamic (γB-A <

1, red), and dynamic (γB-A > 1, blue) simulations; filled data points are determined from simulation, while open circles

are determined by extrapolation. Conditions: CSTR, P of 100 bar, α ~ 0.6, β ~ 102 kJ mol-1, δ ~ 1.4 eV, Yield of B

of 1.0 %, fosc varies, and ΔUB ~ 0.6 eV. For γB-A < 1: T of 100 °C, ΔHovr of 0 kJ mol-1, and asymmetric oscillation

endpoints. For γB-A > 1: T of 150 °C, ΔHovr of -20 kJ mol-1, and symmetric oscillation endpoints.

Page 9: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 8

operating at long times with initially high

concentrations of B (95 mol%) slowly decreases in

concentration until achieving the steady state limit

cycle (red). When starting from a low

concentration of B (10 mol%), the composition

slowly increases until achieving the same limit

cycle (Figure 5B). This steady state solution

(Figure 5C) varies in time in surface coverage and

reactor composition, but the average composition is

fixed in time. As is visually apparent, the

composition approaches the limit cycle only

minimally for each imposed binding energy cycle,

requiring significant computational time to identify

the solution for each parameter set of catalyst and

oscillation conditions.

3.3 Surface Parameters and SuperVolcanoes.

The large number of catalytic parameter

combinations indicates the importance of

categorizing general behaviors of Sabatier-

principle-controlling catalytic reactions. One

strongly-determining parameter is γ, the ratio of the

extent of change of one surface species relative to

another as defined in Equation 9. To understand the

impact of γ on the catalytic reaction, all possible

volcano plots are generated in Figure 6A for 17

variations (0.3 < γB-A < 9.0), while the surface

coverages of A* and B* are depicted for low γB-A

(Figure 6B-6C) and high γB-A (Figure 6D-6E). As

shown, the change in γB-A from above and below

unity dramatically alters the surface coverages and

overall kinetics of the catalytic reaction system.

However, by superimposing all rate volcano plots

on the same panel (Figure 6A), it is apparent that

there exists global behavior; the entire

superimposed set of possible volcano curves are

bounded by a ‘superVolcano’ of extreme limits.

All potential volcano curves are bounded at low

binding energy of B* by the rate of adsorption,

while high bending energy binds the overall rate of

desorption. All variations from the superVolcano

bounds result only from limitations arising from the

surface reaction.

For catalytic systems with high γB-A (>1.0,

blue), the surface product B* lowers in energy

faster than A*, resulting in a decreasing transition

state energy and activation energy by the linear

scaling relationship; for this reason, high gamma

systems exhibit an increasing turnover frequency

with increased binding energy of B* consistent with

a sharp volcano peak depicted in Figure 6A. In

contrast, the low γB-A catalytic systems (γ <1.0, red)

exhibit three regimes. At low binding energy (< -

0.4 eV), the surface is bare, and the overall rate is

controlled by the rate of adsorption; at moderate

binding energies (-0.4 eV < BEB < 0.2 eV), the rate

is controlled by desorption. Even higher binding

energies lead to surface reaction control. The

common point of all curves (0.2 eV relative binding

energy of B*) is equal to the linear scaling

parameter, δ ~ 1.4 eV (offset relative to a binding

energy of B of 1.0 eV).

The surface coverages of the volcano curves of

Figure 6 shift with the binding energy of B*

consistent with the rate-controlling phenomenon.

At low BEB, the surface is bare. For the high γB-A

systems (blue, γ > 1.0), the surface becomes

89.10

89.15

89.20

89.25

89.30

89.35

89.40

89.45

89.50

89.55

89.60

0.4 0.5 0.6 0.7 0.8 0.9 1.0

XB,

Reacto

r C

om

positio

n [

mol

% B

]

θA, Surface Coverage of A [-]

88.10

88.30

88.50

88.70

88.90

89.10

89.30

89.50

0.4 0.5 0.6 0.7 0.8 0.9 1.0

XB,

Reacto

r C

om

positio

n [

mol

% B

]

θA, Surface Coverage of A [-]

Limit

Cycle

Limit

Cycle

Reacto

r C

om

positio

n [

mol%

B]

A B C

Figure 5. Dynamic Forced Oscillation of Surface Binding Energy – Limit Cycles. A simulated catalytic batch

reactor reversibly converting A to B with oscillating binding energy of B* exhibits dynamic variation in surface

coverage and gas-phase composition approaching a limit cycle. (A) Initial conditions above the limit cycle reactor

composition, 95 mol% B. (B) Initial conditions below the limit cycle reactor composition, 10 mol% B. (C) The limit

cycle consists of a three-dimensional stable loop with non-overlapping variations of binding energy of B*, surface

coverage of A, and reactor composition. Catalytic conditions: α ~ 0.6, β ~ 130 kJ mol-1, γB-A ~ 2.0, δ ~ 1.4 eV, f ~

1.0 Hz, ΔUB ~ 0.6 eV, and ΔHrxn ~ 0 kJ/mol.

Page 10: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 9

covered in A* as binding energy increases only to

be completely replaced by B* at the transition point

of +0.2 eV. The opposite behavior exists for low

γB-A systems (red, γ < 1.0); the surface is covered in

A* at high binding energy. The only variation is

the transition observed between high and low

surface coverage of A* and B* at moderate binding

energies.

Four superVolcano plots were created to

compare the impact of changing α, β, and δ on the

superVolcano shapes and kinetic behaviors as

shown in Figure 6A and Figure 7A-7C. The

independent-axis for Figure 6A is zero when the

binding energy of B is 1.0 eV, and the independent-

axis zeros for Figure 7A, 7B and 7C are the

corresponding delta values (0.3, 1.2, and 0.3 eV,

respectively). In the first superVolcano in Figure

6A, low γB-A (< 1.0) systems had higher overall

performance with a maximum TOFB of ~104 s-1 at

relatively weak binding of A and B (0.6 eV for the

binding energy of B). Low γB-A volcano plots

exhibited three distinct regions with varying rate-

limiting steps: (i) adsorption, (ii) surface

reaction/adsorption, and (iii) desorption.

Conversely, high gamma (γ > 1.0) volcano plots

had two primary regions: (i) surface reaction

limited and (ii) desorption-limited regimes. All

volcano plots intersected at a single point,

corresponding to the equivalence point, where the

binding energy of B is (δ + ΔHovr).

Figure 7A demonstrates the changes due to

lowering the value of δ from 1.4 eV to 0.5 eV in

comparison with Figure 6A. At low δ, the high γB-A

curves have higher TOF performance across a wide

span of binding energies, but the maximum TOFB

is lowered to ~102 s-1. In addition, the intermediate

regime of the low γB-A plots moves to the right of

the volcano plot peak at absolute binding energies

of 0.5-1.3 eV, so there now exists a surface

reaction/desorption limited regime. The high γB-A

curves coalesce on the desorption limited line at

~1.3 eV and onwards in Figure 7A, meaning that

these systems are nearly saturated with B* and the

TOFB is solely dependent on the binding energy of

B past this point.

Figure 7B depicts the superVolcano trend for

lower β values as compared with Figure 6A; β was

lowered to 65 kJ/gmol versus 102 kJ/gmol in Figure

6A. The high γB-A curves gain two new rate-limiting

kinetic regimes, with adsorption limited behavior

until an absolute binding energy of 0.4 eV, surface

reaction/adsorption until 0.9 eV, surface

reaction/adsorption from 0.9 eV to 1.4 eV, and

desorption limited behavior from 1.4 eV onwards.

As expected, the overall maximum TOFB increases

to ~105 s-1 since a lower β allows for higher surface

1.0E-08

1.0E-06

1.0E-04

1.0E-02

1.0E+00

1.0E+02

1.0E+04

-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6

Relative Binding Energy of B [eV]

Tu

rnover

Fre

qu

en

cy t

o B

, T

OF

B[s

-1] 102

104

1.0

10-2

10-4

10-6

10-8

0.0

0.2

0.4

0.6

0.8

1.0

-1.0 -0.5 0.0 0.5 1.0

Relative Binding Energy of B [eV]

Su

rface C

ove

rag

e o

f A

[-]

0.0

0.2

0.4

0.6

0.8

1.0

-1.0 -0.5 0.0 0.5 1.0

Relative Binding Energy of B [eV]

Su

rface C

ove

rag

e o

f B

[-]

0.0

0.2

0.4

0.6

0.8

1.0

-1.0 -0.5 0.0 0.5 1.0

Relative Binding Energy of B [eV]

Su

rface C

ove

rag

e o

f A

[-]

0.0

0.2

0.4

0.6

0.8

1.0

-1.0 -0.5 0.0 0.5 1.0

Su

rface C

ove

rag

e o

f B

[-]

Relative Binding Energy of B [eV]

γ > 1.0

γ < 1.0

1.00.90.80.70.60.50.40.3 1.1 1.2 1.3 1.5 1.75 2.0 2.5 4.5 9.0γB

C

D

E

A

Figure 6. Balandin-Sabatier Volcano Curves for Varying γ-Parameter. (A) Turnover frequency to B product.

(B). Surface coverage of A in low γB-A systems. (C) Surface coverage of B in low γB-A systems. (D) Surface coverage

of A in high γB-A systems. €. Surface coverage of B in high γB-A systems. All panels – conditions: YB~1%, P ~ 100

bar, T ~ 150 °C, α ~ 0.6, β ~ 102 kJ mol-1, and δ ~ 1.4 eV.

Page 11: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 10

reaction rates. High γB-A curves coalesce on an

adsorption line from 0.4 eV and lower values of

relative binding energy of B.

Finally, α was raised from the intermediate

value of 0.6 to its maximum value of 1.0, such that

Figure 7C can be readily compared with Figure 7A

since they share all other parameters in common.

The maximum TOFB was relatively unchanged, but

all volcano curves in Figure 7C exhibit steeper

slopes due to the influence of α. The tighter

coalescence of all curves below 0.5 eV shows that

these systems are largely adsorption limited due to

a low δ of 0.5 eV. The desorption lines from 1.3 eV

onwards for low γ systems are shifted to higher

binding energies as compared with Figure 7A. This

is due to the high coverage of A* at binding

energies > 1.3 eV, and this limits the desorption

TOF. In summary, the superVolcano shape and

kinetic behavior is highly sensitive to both the

transition state BEP relationship parameters and the

binding energy equivalence point, δ. Since these

parameters are a strong function of the reaction

chemistry and metal crystal face/structure, practical

application of dynamic catalysis will be guided by

a combination of catalyst design and external

stimulus selection to achieve beneficial catalytic

reaction control.

3.4 Low Gamma Catalytic Kinetics. The

dynamic kinetics of high gamma (γB-A ~ 2.0)

catalytic systems were previously evaluated and

shown to exhibit significant rate enhancement

(1,000-10,000x above the Sabatier maximum) in

the resonant frequency range[24]. Under optimal

conditions, the surface coverage of high γ

chemistries oscillate between high coverage in A*

and high coverage in B*, ultimately yielding a

turnover frequency about equal to the imposed

surface oscillation frequency. As these dynamic

catalytic systems modulated into the strong binding

state, the heat of the surface reaction was at its most

negative condition leading to the lowest energy

transition state; this permitted A* to readily convert

to B*. When the surface binding energy then

converted to weaker overall binding, the binding

energy of B* decreased thereby decreasing the

desorption barrier. Additionally, the high energy

binding state inhibited the reverse surface reaction

(B* to A*) due to a high transition state energy.

This behavior held for multiple waveform types

(e.g. square, sinusoidal) over a range of linear

transition state scaling relationships (0 < α < 1.0),

making these types of reactions amenable to

dynamic rate enhancement.

Low gamma (γB-A <1.0) catalytic systems

exhibit similar behavior with the variation of

surface intermediate binding, but conditions

leading to surface resonance remain to be

identified. As depicted in Figure 8A, dynamic

catalysis can lead to performance above the

Sabatier maximum rate for low gamma systems. In

this example, catalytic conversion of A-to-B occurs

in a continuous-flow catalytic reactor at fixed

temperature and pressure, and the system switches

from static operation to five different dynamic

frequencies for comparison (f ~ 0.001, 0.035, 1.0,

10, and 1000 Hz, ΔUB ~ 0.93 eV). Oscillation

frequencies above the Sabatier maximum rate

(~0.036 s-1) lead to high average turnover frequency

1.E-09

1.E-07

1.E-05

1.E-03

1.E-01

1.E+01

1.E+03

-1.0 -0.5 0.0 0.5 1.0

Turn

ove

r F

reque

ncy [

s-1

]

Relative Binding Energy of B [eV]

1.E-13

1.E-10

1.E-07

1.E-04

1.E-01

1.E+02

1.E+05

-1.5 -1.0 -0.5 0.0 0.5 1.0

Turn

ove

r F

reque

ncy [

s-1

]

Relative Binding Energy of B [eV]

1.E-11

1.E-09

1.E-07

1.E-05

1.E-03

1.E-01

1.E+01

1.E+03

1.E+05

-1.0 -0.5 0.0 0.5 1.0

Turn

ove

r F

reque

ncy [

s-1

]

Relative Binding Energy of B [eV]

1.00.90.80.70.60.50.40.3 1.1 1.2 1.3 1.5 1.75 2.0 2.5 4.5 9.0γ

10-7

10-9

10-5

10-3

10-1

101

103

10-7

10-10

10-13

10-4

10-1

102

105

10-9

10-7

10-5

10-3

10-1

101

103

105

10-11

B CA α ~ 0.6

β ~ 102 kJ/mol

δ ~ 0.5 eV

α ~ 0.6

β ~ 65 kJ/mol

δ ~ 1.4 eV

α ~ 1.0

β ~ 102 kJ/mol

δ ~ 0.5 eV

Figure 7. Parameter Variations Catalytic Reactions Depicted as Balandin-Sabatier SuperVolcanoes. A

catalytic flow reactor reversibly converts A to B at 150 °C, 100 bar of pure feed A, and 1% yield of B. In each case,

the heat of reaction was exothermic, ΔHrxn ~ -20 kJ/mol with varying gamma, 0.3 < γB-A < 9.0. Surface parameters

varying between cases included: (A) Alpha of 0.6, beta of 102 kJ mol-1, and delta of 0.5 eV. (B) Alpha of 0.6, beta

of 65 kJ mol-1, and delta of 1.4 eV. (C) Alpha of 1.0, beta of 102 kJ mol-1, and delta of 0.5 eV.

Page 12: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 11

up to 25 s-1. The depicted instantaneous TOFB for

each frequency shows flipping between high TOFB

and low TOFB states due to the large change in

surface coverages of A* and B*.

The low gamma system (γB-A ~ 0.5) of Figure

8B exhibits a broad range of resonance frequencies.

For an oscillation amplitude of 0.93 eV, the TOFB

frequency response is shown in Figure 8B for a

range of imposed square wave oscillation

frequencies (10-6 < f < 1012) in the binding energy

of B*. From this plot, four corner frequencies can

be determined as the transitions in slope; the

resonant frequency range extends from about 102 to

107 Hz. Frequencies below 10-4 Hz yield a TOF of

~0.02 s-1, which is the average of the steady state

TOF at two oscillation endpoints, while frequencies

above 1011 Hz yield the TOF from the starting

condition (i.e. the volcano plot peak).

Figures 8A and 8B can be directly compared to

the performance of the high gamma system (γB-A ~

2.0) depicted in our previous publication[24]. At

0.001 Hz, the TOFB response in Figure 8A exhibits

an overshoot consistent with the flipping of the

surface binding energies and off-loading of the

surface. At intermediate frequencies (0.035, 1.0,

and 10 Hz) in Figure 8A, the TOFB response has

sharp features due to rapid uptake of gas-phase

species. In the resonant region (1000 Hz), a

significant number of oscillations are required for

the system to achieve steady state consistent with a

gradual change in the surface coverages; for this

condition, there exists an initial overshoot of the

average TOFB above the final steady state value.

This example indicates that the catalytic response

of both low and high gamma systems, in this case

γB-A ~ 0.5 and γB-A ~ 2.0, can exhibit similar general

reaction behavior, with the caveat that any

particular dynamic catalytic system must be

evaluated to identify the conditions that permit

resonance and rate enhancement. This key

observation is relevant to general applicability of

dynamics, since forward and reverse directions of

any reaction will exhibit inverse values of gamma;

this indicates that at least some catalytic systems

could be dynamically promoted in either direction

(forward or reverse) depending on the selected

parameters of the imposed surface oscillation.

While both the forward and reverse directions

of a reaction can be promoted dynamically, the

strategies for selecting the parameters of the

imposed surface binding energy oscillation

between systems are indeed different. For the low

gamma system (γB-A ~ 0.5) depicted in Figure 9A,

the Sabatier curve is depicted in black with

extended dashed lines above the Sabatier peak. The

oscillation amplitude is depicted as a red bar with

two endpoints: (i) the strong binding energy

maximum is located at the position of the volcano

peak, and (ii) the weak binding energy minimum is

permitted to vary. The average turnover frequency

to B is depicted in the heatmap of Figure 9B with

variable square wave surface frequency and tunable

1.E-03

1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

Tu

rno

ve

r F

req

ue

nc

y t

o B

[s

-1]

Time on Stream [Arbitrary Units]

10-3

10-1

1

10+1

10-2

10+2A

0.001

0.01

0.1

1

10

100

1E-06 1E-02 1E+02 1E+06 1E+10

Tim

e A

ve

rag

ed

TO

FB

[s-1

]

Oscillation Frequency [Hz]

B

1000 Hz

10 Hz

1.0 Hz

0.001 Hz

0.035 Hz

Figure 8. Kinetics of Low Gamma (γ ~ 0.5) Continuous Flow Catalytic Reactor. (A) A continuous flow reactor

operating under static conditions exhibits TOFB ~ 0.036 s-1; implementation of catalyst dynamics at 0.001 Hz (red),

0.035 Hz (green), 1.0 Hz (light blue), 10 Hz (dark blue) and 1000 Hz (purple) varies the instantaneous TOFB. (B)

Continuous variation of the catalyst binding energy over varying frequencies (10-6 < f < 1011 s-1) reveals a band of

resonance frequencies highlighted in purple. Conditions: T ~ 100 °C, P ~ 100 bar feed of A, YB ~ 1%, ΔHrxn ~ 0 kJ

mol-1, α ~ 0.6, β ~ 102 kJ mol-1, γB-A ~ 0.5, δ ~ 1.4 eV, ΔUB ~ 0.93 eV, [0.10 eV < UB < 1.03 eV]

Page 13: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 12

minimum binding energy of B. Until the oscillation

amplitude low binding energy endpoint (left) is less

than -0.4 eV relative binding energy of B, the time

averaged turnover frequency is negligible, and the

A → B reaction is inhibited by dynamic catalysis.

Above the -0.4-eV transition, the average turnover

frequency to B is about constant and oscillation

frequencies above 103 Hz lead to a high rate

performance of 25 s-1 for a total rate enhancement

of ~715x.

The importance of the transition at about -0.4

eV relative binding energy of B is indicative of the

mechanism leading to dynamic enhancement. As

shown in the surface coverage plot of Figure 6C,

this transition is associated with the shift from a

clean surface (θ* ~ 1) to a surface covered in B* (θB

~ 1). When the oscillation amplitude extends into

the low binding energy range associated with a

clean surface, then B* is removed from the surface

producing a dramatic reaction rate enhancement.

The significance of transitioning between a

surface covered in reaction surface species and a

clean surface for low gamma systems under

dynamic conditions has implications for selection

of catalyst materials. In general, the binding

energies of A* and B* need to be sufficiently weak

that all chemical species can be desorbed during the

weaker of the surface oscillation states; in this case

δ ~ 1.4 eV exhibited this capability. However, as

the binding energy of B* varied with the imposed

surface oscillation, a new limitation exists; the heat

of adsorption of both B* and A* must be less than

or equal to zero. For low gamma systems, small

variations in the binding energy of B* produce large

variations in the binding energy of A* such that it

readily approaches zero enthalpy of adsorption.

This scenario can be interpreted by the gamma-

delta plot of Figure 2B; as binding energies weaken;

one surface species will achieve negligible

adsorption enthalpy before the other one. In this

scenario, the surface species with negligible

adsorption enthalpy will remain in that state as the

other surface species continues to change with the

imposed oscillation.

The physical restriction of non-repulsive

surfaces for adsorption alters the enthalpy of the

surface reaction and the surface transition state

energy by association. The implication is apparent

in the TOFB heat map of Figure 9C, where the TOFB

for the surface reaction elementary step is presented

(color) as a function of the surface coverage of A

(θA) and the relative binding energy of B [eV]. At

low relative binding energies of B, the system

exhibits a gradual change in the TOFB before a

sharp transition above -0.7 eV; this transition is the

condition where the binding energy of A

approaches zero. Figure 9C also reveals part of the

origin for fast dynamic TOFB; at high surface

coverage of A, the surface reaction can achieve

TOFB of ~102 s-1. These unique behaviors indicate

that three-order-of-magnitude enhancement in rate

is achievable for low gamma systems, but selection

1.E-08

1.E-06

1.E-04

1.E-02

1.E+00

1.E+02

1.E+04

1.E+06

-1.0 -0.5 0.0 0.5

Turn

ove

r fr

equ

ency

(1

/s)

Relative binding energy of B (eV)Relative Binding Energy of B [eV]

Tu

rnover

Fre

qu

en

cy [

s-1

]

1

106

104

102

10-2

10-4

10-6

10-8 1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

1.E+05

-1.0 -0.8 -0.6 -0.4 -0.2 0.0

Osc

illat

ion

fre

qu

ency

(H

z)

Relative binding energy of B (eV)Relative Binding Energy of B, Left Endpoint [eV]

Oscill

ation

Fre

qu

en

cy [

s-1

]

FAST SLOW

Turnover Frequency to B, TOFB [s-1]

105

104

103

102

101

1

10-1

10-2

10+1 1 10-1 10-2

-1.40

-1.18

-0.96

-0.74

-0.52

-0.30

-0.08

1.E-10 1.E-08 1.E-06 1.E-04 1.E-02 1.E+00

Re

lati

ve

Bin

din

g E

ne

rgy o

f B

[e

V]

Surface Coverage of A, θA [-]Surface Coverage of A, θA [-]

Rela

tive B

ind

ing

En

erg

y o

f B

[eV

]

10-10 10-8 10-6 10-4 10-2 1

Turnover Frequency to B, TOFB [s-1]

10+2 10+1 1 10-1 10-2 10-3A B C

Figure 9. Catalytic Dynamics in a Flowing Stirred Tank Reactor at Differential Conversion with Variable

Amplitude at Low Gamma. (A) Volcano kinetics of reversible A to B with variable amplitude. (B) Average

turnover frequency to B (color distribution) as a function of square wave oscillation frequency and the amplitude low

energy endpoint; the high energy endpoint is fixed at +1.03 eV. (C) Maximum possible surface reaction rate (color

distribution) for A* reversibly converting to B* as a function of variable surface coverage and binding energies;

negligible reverse reaction occurred under these conditions. Conditions: T~ 100 °C, P~100 bar A feed, YB ~ 1%,

ΔHrxn ~ 0 kJ mol-1, α ~ 0.6, β ~ 102 kJ mol-1, γB-A ~ 0.5, δ ~ 1.4 eV, and ΔUB ~ variable.

Page 14: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 13

of the dynamic oscillation of surface binding

energies must be carefully selected unique to each

system.

3.5 Dynamic Steady State. As stated in the

introduction, a grand challenge in catalysis and

reaction engineering targets high yields of a desired

product in the constraint of thermodynamic

equilibrium and the mass action of Le Chatelier’s

principle. As shown in Figure 10A, reaction

systems can be thought of as a thermodynamic

trajectory between the reactant and product. Under

thermodynamic control, the reaction system

performance will approach the product composition

where the overall free energy, or chemical

potential, is minimized. Since each molecule has

significantly different free energy due to varying

chemical potentials, bonding, and functional

groups, this free energy landscape is a strong

function of the reaction chemistry.

The importance of controlling extent of

reaction has led to the strategy depicted in Figure

10B; application of external work to one or both of

the two thermodynamic states perturbs the free

energy landscape in favor of a new equilibrium

extent of conversion. For example, work via

compression of reactant gases in chemistries such

as hydrogenation of hydrocarbons alter the free

energy landscape in favor of saturated bonds,

requiring high pressure hydrogen as a co-reactant.

Dynamic oscillation of surface binding energies

potentially permits work to be imparted directly to

the surface reaction (Figure 10C). Since work is

applied directly to the surface reaction, the free

energy landscape between A and B is modified

without changing the free energy of the gas phase

species. A system of this type is depicted in Figure

10D, with the minimum energy offset from

equilibrium by ΔΔG. Thermodynamic work as a

‘catalytic molecular pump’ moving surface

adsorbates from strong to weak binding could

therefore permit reactor systems that achieve high

catalytic conversion at thermodynamically

unfavorable gas-phase conditions.

To assess catalytic reactor performance and

extent of reaction of dynamic catalytic systems,

simulations were conducted using a batch reactor

charged with 100 bar of initial gas at varying

composition of A and B with a fixed amount of

catalyst with dynamically varying surface binding

energies. Figure 11 depicts the time-resolved gas-

phase composition within the catalytic batch reactor

(100 bar, 250 °C) with seven different initial

compositions of A and B. Dynamic surface binding

energy of the catalyst was applied for 200 seconds

of square wave of amplitude 0.5 eV and 1000 Hz,

after which the system was converted to static

catalysis (in this case at the volcano plot peak

binding energy). Regardless of the initial gas-

phase composition of the reactor, the system

approached a dynamic steady state of 70 mol% B in

the gas phase. While the small oscillatory behavior

is difficult to observe in Figure 11, oscillation in the

instantaneous TOFB was observed upon zooming in

on any of the data sets. Once the catalyst was

converted from dynamic to static operation at 200

s, the system returned to thermodynamic

equilibrium at 40 mol % B (ΔHovr of +2 kJ/gmol),

indicating that deviation from equilibrium resulted

from dynamic ‘catalytic pumping’ on the catalyst

surface.

Gib

bs F

ree E

nerg

y, G

Conversion

ΔGEQ

ΔΔ

G

XEQ

XSS

AG

ibbs F

ree E

nerg

y, G

Conversion

ΔGEQ

XEQ

XEQ

ΔΔ

G

D

State A State B

Reaction

Path

State A State BReaction

Path

WORKWORK

B

C

Figure 10. (A) Gibbs free energy landscape for a reaction with differing overall heats and entropy of reaction in state

B, ΔΔG. (B) Work applied to reactant state, B, alters the overall Gibbs free energy of reaction and the extent of

conversion at equilibrium. (C) Work applied in the reaction path between reactant A and product B alters the Gibbs

free energy minimum and extent of conversion at steady state. (D) General reaction system Gibbs free energy with

respect to extent of conversion with added work, ΔΔG, and minima identified as equilibrium points.

Page 15: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 14

Figure 11. Dynamic Catalytic Reaction to Steady

State Different from Equilibrium - High γB-A

Condition. The reversible reaction of A to B undergoes

dynamic catalytic conversion in a batch reactor at 250

°C, 100 bar total pressure, and a square waveform at

1000 Hz and 0.5 eV amplitude. Variable initial

concentrations of component A. Surface chemistry

parameters, α ~ 0.6, β ~ 135 kJ mol-1, γB-A ~ 2.0, δ ~ 1.0

eV, and ΔHrxn ~ +2.0 kJ mol-1.

3.6 Oscillatory Steady State of High γ

Catalytic Systems. The steady state performance

of a representative high gamma catalytic system

(γB-A ~ 2.0, δ ~ 1.4 eV) was assessed using a batch

reactor model. Sabatier volcano plots for this

system (α ~ 0.6, β ~ 102 kJ/mol) are shown in

Figure 12A, with reactor temperatures varying

between 150-250 oC. Qualitatively, temperature

does not change the shape and the rate-limiting

kinetic regimes of the volcano plots. Similarly, the

associated surface coverages for the considered

system was plotted in Figure 12B for both A* and

B*. B* dominated the surface at high relative

binding energies while A* dominated at weaker

binding energies. Each set of curves crossed over at

the delta point, which is the equivalence point

where A* and B* have the same energy. As

temperature increased, the onset of surface vacancy

shifted to stronger binding energies, and the

remaining coverage of A* decreased faster with

higher temperature (not shown, BE < -1.0 eV).

The extent of conversion at oscillatory steady

state of the considered system of Figure 12 was

evaluated as a function of applied square wave

frequency f, waveform amplitude, and waveform

amplitude position. As depicted in Figure 12C at

150 °C, dynamic oscillation of the surface binding

energy of B at a fixed amplitude can occur over a

range of oscillation endpoints. The position of the

strong binding energy of B endpoint was the

independent variable in the heat maps of Figures

12D (ΔUA ~ 0.5 eV), 12E (ΔUA ~ 1.0 eV), and 12F

(ΔUA ~ 1.5 eV), where the oscillatory steady state

TOFB was determined via simulation.

In the simulations of Figure 12D-12F, the heat

and entropy of reaction were zero, indicating that

the system should achieve only 50% composition

of both A and B in the gas phase (green) at

thermodynamic equilibrium. However, the broad

color range associated with the concentration at

oscillatory steady state indicates that there exist

catalytic conditions yielding substantial shift away

from equilibrium as high as >99% yield of B for all

oscillation frequencies above 1.0 Hz and oscillation

endpoints above about 0.2 eV. This region of high

conversion to B at steady state extends over a broad

range that ends at lower oscillation endpoints

before transitioning into a region favorable to

forming the reactant species. Frequencies this slow

and even as high as 0.1 Hz are achievable for all

conceivable experimental methods of imposing

oscillatory surface binding energy of adsorbates

including oscillating electric fields and surface

stress/strain.

The higher amplitude heat map plots of TOFB

at ΔUA ~ 1.0 eV (Figure 12E) and ΔUA ~ 1.5 eV

(Figure 12F) show enhanced performance at lower

frequency (about 10-2 s-1), because the amplitude

oscillation endpoints move further and further away

from the volcano plot peak. Moreover, there exists

a sharp transition between the forward and

backwards reaction at 0.4 eV relative binding

energy, which exactly matches the absolute value

of delta, δ ~ 1.4 eV. High conversion to either A or

B (>99%) is achievable at these higher amplitudes

requiring only the selective application of surface

binding oscillation at the relevant frequency and

oscillation amplitude endpoints.

3.7 Oscillatory Steady State of Low γB-A

Systems. Low gamma catalytic systems introduced

in section 3.4 exhibited dramatic rate enhancement

under dynamic operation (Figure 8) but with more

complex kinetic behaviors as compared with high

gamma systems in CSTR simulations. To assess the

potential for controlling the extent of reaction of

low gamma catalysis, a simulation with parameters

Page 16: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 15

identical to the volcano plot of Figure 12A (except

for a γB-A of 0.5) was conducted in a batch reactor

with a square wave in surface binding energy

oscillation. As depicted in Figure 13A-13C, three

oscillation amplitudes of ΔUB of 0.5, 1.0, and 1.5

eV were evaluated as a function of oscillation

endpoints and frequency. Again, the simulated

system reaction equilibrium was 50% composition

(green in Figure 13) of both A and B.

For an oscillation amplitude of ΔUB ~ 0.5 eV in

Figure 13A, there exists a small region (yellow)

where the forward A → B reaction is promoted

between relative binding energies of B of 0-0.2 eV

and about 102 Hz. However, above about 0.15 eV

of relative binding energy of B exists a large region

(dark blue) overwhelmingly favorable to the

reverse reaction (formation of A). Higher

oscillation amplitudes in low gamma γB-A catalytic

systems again lead to high conversion at oscillatory

steady state, in this case for both reaction directions

as yields > 90 % B are observed while other

oscillation endpoints yield > 99 % A. The two

regions are split by the delta point (δ ~ 1.4 eV, 0.4

relative binding energy of B). The implication is

that this system can be tuned to produce an outlet

stream with nearly completely controllable

composition at high reactor residence times, just by

varying the oscillation endpoints at frequencies

between 1-10 Hz. The high and low conversion

regimes exist as low as 0.001 Hz making them

accessible to slowly oscillating systems. This

behavior is nearly identical to the high gamma

system in Figure 12; as the only difference between

Figure 12 and 13 being the inverse values of gamma

(γB-A of 2.0 and 0.5, respectively), this inverse

mirrored behavior is expected. Both low and high

gamma systems provide versatility in conversion as

nearly pure product streams of A or B can be

produced by merely changing the dynamic catalysis

oscillation endpoints relative to the catalytic

chemistry delta point.

3.8 Mechanism of Catalytic Molecular

Pumping. The tunable directionality (forward

versus reverse) of catalytic molecular pumping

observed in Figures 12 and 13 derives from the

mechanism of molecular movement through the

oscillating energy profiles on the catalytic surface.

As shown earlier, catalytic systems can be driven in

the forward or reverse directions by selection of the

catalytic material or stimulus method (affecting α,

0.0

0.2

0.4

0.6

0.8

1.0

-1.0 -0.5 0.0 0.5 1.0

Surf

ace

cove

rage

(u

nit

less

)

Relative binding energy of B (eV)

1.E-08

1.E-06

1.E-04

1.E-02

1.E+00

1.E+02

-1.0 -0.5 0.0 0.5 1.0

Turn

ove

r fr

equ

ency

(1/s

)

Relative binding energy of B (eV)

Binding Energy of B* [eV]

-1.0 0.5 1.00-0.5

Tu

rnover

Fre

qu

en

cy t

o B

[s

-1]

10-2

Binding Energy of B* [eV]

-1.0 0.5 1.00-0.5

Su

rface C

overa

ge [

-]

1.00

0.80

0.60

0.40

0.20

0

250 C

200 C

150 C

θA

θB

1.00E-06

1.00E-05

1.00E-04

1.00E-03

1.00E-02

1.00E-01

1.00E+00

1.00E+01

1.00E+02

1.00E+03

1.00E+04

1.00E+05

1.00E+06

-1.00 -0.50 0.00 0.50 1.00

Turn

ove

r Fr

equ

ency

to

B [s

-1]

Binding energy shift of B* (eV)

Binding Energy of B* [eV]

-1.00 0.50 1.000-0.50

Tu

rnover

Fre

qu

en

cy t

o B

[s

-1]

P1

P2P3

P4P5

106

104

102

100

10-2

10-4

A B C

1E-06

1E-05

1E-04

1E-03

1E-02

1E-01

1E+00

1E+01

1E+02

0.0 0.2 0.4 0.6 0.8 1.0

Osc

illat

ion

freq

uen

cy (H

z)

Oscillation endpoint (eV)

102

100

10-1

10-2

10-3

10-4

10-5

10-6

Oscill

ation

Fre

qu

en

cy [

s-1

]

Oscillation Endpoint, B [eV]

0 0.40 1.60 2.001.200.80

E

101

ΔUA = 1.0 eV1E-06

1E-05

1E-04

1E-03

1E-02

1E-01

1E+00

1E+01

1E+02

0.0 0.5 1.0 1.5 2.0 2.5 3.0

Osc

illa

tio

n fr

eq

ue

ncy

(Hz)

Oscillation endpoint (eV)

Con

vers

ion

of A

[%]

100

90

80

70

60

50

40

30

20

10

0

F 102

100

10-1

10-2

10-3

10-4

10-5

10-6

Oscill

ation

Fre

qu

en

cy [

s-1

] 101

Oscillation Endpoint, B [eV]

0 0.5 2.5 3.01.51.0 2.0

ΔUA = 1.5 eV1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

0.0 0.2 0.4 0.6 0.8 1.0

Turn

ove

r fre

qu

ency

(1/s

)

Relative binding energy of B (eV)

101

100

10-1

10-2

10-3

10-4

10-5

10-6

Oscill

ation

Fre

qu

en

cy [

s-1

]

Oscillation Endpoint, B [eV]

-0.2 0 0.6 0.80.40.2

D102

ΔUA = 0.5 eV

10-4

10-6

10-8

1

102

Figure 12. Dynamic Catalytic Conversion of A to B for High Gamma (γ ~ 2.0, δ ~ 1.4 eV). (A) Turnover

frequency to product chemical B from reactant A for α ~ 0.6, β ~ 102 kJ/mol, and YB~1% for 150, 200, and 250

°C. (B) Surface coverages of A* (θA) and B* (θB) for α ~ 0.6 and YB~1% for 150, 200, and 250 °C. (C) Varying

oscillation endpoints for ΔUB ~ 0.8 eV. (D/E/F) Oscillatory steady state conversion of A (0 < XA < 100%) at 150 °C

and 100 bar for varying applied surface oscillation frequency and oscillation endpoint with oscillation amplitudes

ΔUA of 0.5, 1.0, and 1.5 eV

Page 17: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 16

β, γ, and δ) but also by the selection of imposed

surface oscillation including the frequency,

amplitude, and amplitude endpoints. The

relationship between selected parameters and

catalytic pumping directionality is depicted in

Figure 14 for the system previously described in

Figure 13 with a low gamma (ΔHovr ~ 0 kJ gmol-1, α

~ 0.6, β ~ 102 kJ gmol-1, γB-A ~ 0.5, and δ ~ 1.4 eV).

Two conditions were selected; Figure 14A-B

correspond to an oscillation amplitude promoting

the reverse reaction, while Figure 14C-D has an

amplitude promoting the forward reaction. In each

energy diagram (Figure 14A, 14C), the amplitude

minimum (Umin, blue) and maximum (Umax, red) are

depicted along with the relative binding energy of

B* corresponding to the Sabatier peak under static

conditions (purple, dashed) and the binding energy

associated with delta (black, δ ~ 1.4 eV).

The directionality of the dynamic catalytic

system is visually apparent from the values of the

enthalpy of A* and B* at the minimum amplitude

(Umin). The reverse reaction (B-to-A) is favored

when A* is lower in energy than B* at Umin in

Figure 14A, while the forward reaction (A-to-B) is

favored when B* is lower in energy than A* at Umin

in Figure 14C. By this interpretation, the lower

energy state Umin serves as the condition whereby

the surface reaction proceeds to accumulate A* or

B* on the surface; the higher energy state Umax then

serves to push this surface reaction product into the

gas phase. In the low energy state Umin, the

transition state enthalpy must be lower than the

desorption enthalpy to permit the surface reaction

to proceed. This general behavior is depicted in the

schemes for the reverse reaction (Figure 14B) and

forward reaction (Figure 14D).

Interpreting general reaction systems for

potential catalytic molecular pumping relies on the

determination of the reaction energy profile under

different oscillation conditions with respect to the

parameter delta, δ. These intermediate and

transition state energies can now be determined

computationally for almost any catalytic system[66].

As defined earlier, the quantity delta δ identifies the

adsorption enthalpy whereby both A* and B* have

equal binding energy with the surface. Delta

therefore serves as a separation point between the

forward and reverse directionality of catalytic

molecular pumping; for any system with a non-

unity gamma, the surface species of lowest energy

(A* or B*) will differ for systems operating on

either side of delta. This precise transition in

directionality was observed in Figures 12 and 13,

where the shift from high to low conversion was

demarcated by the relative binding energy of B*

equaling delta.

3.9 Efficiency of Dynamic Catalysis at

Oscillatory Steady State. Returning to the concept

of imparting work during catalysis as described in

Figure 10C, the observed free energy change for the

reaction (ΔΔG) is related to the applied oscillation

amplitude (ΔUi). In mechanical applications, this

relationship is typically expressed as an efficiency,

which compares the applied work (i.e. energy units

after integrating over time) to the resulting change

in system enthalpy, internal energy, or another

indicator such as temperature. Here, the efficiency

of the dynamic catalytic system is defined as,

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

0 0.1 0.2 0.3 0.4 0.5

Oscill

atio

n F

reque

ncy [

Hz]

Relative Binding Energy of B [eV]

Ste

ady-S

tate

Co

nve

rsio

n [%

]

10-1

10-2

10-3

10-4

10-5

10-6

1

101

102

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

1.E+01

0.0 0.2 0.4 0.6 0.8 1.0

Oscill

atio

n F

reque

ncy [

Hz]

Relative Binding Energy of B [eV]

10-5

10-4

10-3

10-6

10-2

10-1

1

101

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

1.E+01

0.00 0.25 0.50 0.75 1.00 1.25 1.50

Oscill

atio

n F

reque

ncy [

Hz]

Relative Binding Energy of B [eV]

10-1

1

101

10-2

10-3

10-4

10-5

10-6

A B C

ΔUB = 0.5 eV ΔUB = 1.0 eV ΔUB = 1.5 eV

Figure 13. Dynamic Catalytic Conversion of A to B for Low Gamma (γB-A ~ 0.5, δ ~ 1.4 eV). Steady-state

average conversion of A-to-B in a batch reactor (equilibrium at 50% - green) for varying square wave amplitude high

binding energy state [eV] and oscillation frequency [Hz] for fixed amplitudes ΔUB of: (A) 0.5 eV, (B) 1.0 eV, and

(C) 1.5 eV. Batch reactor conditions: 150 °C, initial reactor composition of 100 bar pure A, ΔHrxn ~ 0 kJ mol-1, α ~

0.6, β ~ 102 kJ mol-1, γB-A ~ 0.5, and δ ~ 1.4 eV.

Page 18: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 17

휀 =|𝛥𝛥𝐺|

𝛥𝑈𝑖=

𝐹𝑟𝑒𝑒 𝐸𝑛𝑒𝑟𝑔𝑦 𝐷𝑒𝑣𝑖𝑎𝑡𝑖𝑜𝑛

𝐴𝑝𝑝𝑙𝑖𝑒𝑑 𝑂𝑠𝑐𝑖𝑙𝑙𝑎𝑡𝑖𝑜𝑛 𝐴𝑚𝑝𝑙𝑖𝑡𝑢𝑑𝑒

(10)

The efficiency is the absolute value of the free

energy deviation from equilibrium to account for

both the forward and reverse reaction promotion at

oscillatory steady state relative to the applied

oscillation amplitude. For consistency with the

mechanism of dynamic catalysis of Figure 14, the

amplitude in the denominator of the efficiency

(ΔUA or ΔUB) was selected based on the

directionality of the promoted reaction; ΔUB was

selected for the forward reaction, and ΔUA was

selected for the reverse reaction.

Representative high and low gamma systems

were analyzed to assess the efficiency as a function

of oscillation amplitude, waveform, and endpoint

selection. As demonstrated in a previous

publication for a continuous flow reactor under

dynamic catalysis operation[24], oscillation

waveform shape has a significant effect on dynamic

catalysis rate promotion. In the CSTR for a high

gamma system (γB-A ~ 2.0), the square waveform

exhibited the highest TOFB followed by sinusoidal

(2-3x slower) and then triangle and sawtooth

waveforms (4-6x slower). As shown in Figure 15,

these waveform types were applied to the batch

reactor system, and the resulting steady state

composition was compared with the

thermodynamic equilibrium composition (50% A

and B in all cases). The ΔΔG was computed as the

difference in the apparent free energy of the

-250

-200

-150

-100

-50

0

50

100

En

tha

lpy [

kJ

mo

l-1]

Umin

Umax

δ

Static

Peak

A* B*

Reverse

A

B

-175

-125

-75

-25

25

75

125

En

tha

lpy [

kJ

mo

l-1]

A* B*

δ

Umin

Umax

Static

Peak

Forward

C

D

Umax

Umin

Umax

Umin

Reverse Forward

Figure 14. Mechanism of Catalytic Molecular Pumping and Rate Enhancement. Energy diagrams for catalytic

unimolecular reaction system with reaction chemistry defined as: ΔHovr ~ 0 kJ gmol-1, α ~ 0.6, β ~ 102 kJ gmol-1, γB-

A ~ 0.5, and δ ~ 1.4 eV. (A) Oscillation amplitude of 1.0 eV with oscillation endpoints of B* of Umin ~ 0.6 eV and

Umax ~ 1.6 eV. (B) For the selected oscillation amplitude, the catalytic molecular pump moves molecules of B in

green through adsorption to B* and then reaction to A*; subsequent oscillation desorbs A* to A. (C) Oscillation

amplitude of 1.0 eV with oscillation endpoints of B* of Umin ~ 0.2 eV and Umax ~ 1.2 eV. (D) For the selected

oscillation amplitude, the catalytic molecular pump moves molecules of A in orange through adsorption to A* and

then reaction to B*; subsequent oscillation moves molecules of B* to product B in the gas phase.

Page 19: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 18

reaction at oscillatory steady state from the actual

equilibrium (i.e., 50% A and 50% B).

Figure 15A shows the efficiency of the applied

waveforms as a function of imposed amplitude at a

constant frequency of 10 Hz. Initially, near zero

amplitude, the efficiency is low for all waveforms

as shown by the shallow slope of the plot. Once the

oscillation amplitude achieves 10-20 kJ mol-1 (i.e.

0.1-0.2 eV), the relationship between ΔΔG and ΔU

is effectively described by a straight line. The slope

of this line is the system efficiency during dynamic

catalysis since it is the ratio of two energy units. The

maximum efficiency observed was ~40 % for the

square wave with the sinusoidal and triangular

wave being slightly less effective at 34% and 30%,

respectively.

Due to the importance of oscillation amplitude

endpoints in low gamma γB-A catalytic systems, the

efficiency of low gamma systems was evaluated in

Figure 15B as a function of both oscillation

amplitude and oscillation amplitude endpoint in a

flow reactor. As observed in the low gamma batch

reactor results in Figure 13A-13C, the forward A →

B reaction (yellow-to-red) was only enhanced at

oscillation endpoints up to the delta point. Once

stronger binding energies were imposed on the

system, the B → A reaction (blue) was highly

favorable due to the dominant surface coverage of

A*. The highest observed efficiency here was ~16

%, specifically for the B → A reaction.

A parity region was observed in Figure 15B

that runs diagonally along the blue color-coded B

→ A region with a slope of about unity. This region

traces points where the difference between the

oscillation endpoint and amplitude equals delta.

Once this difference was less than delta (i.e. the

lower right quadrant of the heat map), the system

performance rapidly declined to 0% efficiency

(green). This inefficiency resulted from the surface

coverage which needs to be turned over

significantly to enhance reactor performance, and

the delta point determines the crossover between

A* and B* covered surfaces. In summary, the

system efficiency for high gamma and low gamma

cases is a strong function of oscillation amplitude,

frequency, and endpoint selection but waveform

shape also has a minor effect.

3.10 Catalytic Molecular Pump and

Molecular Machines. The design of molecules

with dynamic functionality has been extensively

described as ‘molecular machines’[67] with both

synthetic and biological examples[68,69]. While

‘switchable catalysts’ are capable of turning on and

off as needed[70], other molecular machines and

molecular devices are capable of implementing

motion including molecular motors and

pumps[68,71,72]. For example, an artificial molecular

pump is capable of moving molecules such as

charged rings up a concentration gradient[73],

conceptually similar to nature’s capability for

0.0

0.3

0.6

0.9

1.2

1.5

0.0 0.3 0.6 0.9 1.2 1.5

Oscill

atio

n A

mplit

ude

[e

V]

Oscillation Endpoint [eV]

Surf

ace

Oscill

atio

n E

ffic

iency,

ε[%

]

Fo

rwa

rdR

eve

rse

20

15

10

5

0

5

10

15

20-40

-35

-30

-25

-20

-15

-10

-5

0

0 20 40 60 80 100

Amplitude [kJ mol-1]

De

lta D

elta

Gib

bs F

ree

Ene

rgy,

ΔΔ

G [

kJ m

ol-1

]

Square (40.1%)

Sine (34.1%)

Triangle (30.2%)

A B

Oscillation Endpoint [eV]

Figure 15. Conversion Efficiency of A-to-B Reaction on Dynamic Catalytic Surfaces. (A) Square, sinusoidal,

or triangle wave with varying amplitude operating at 10 Hz (α ~ 0.6, β ~ 135 kJ mol-1, γ ~ 2.0, δ ~ 1.0 kJ mol-1, and

ΔHovr ~ 0 kJ mol-1) on a catalyst in a batch reactor at oscillatory steady state offset from equilibrium by delta delta

Gibbs free energy, ΔΔG. The efficiency is defined as the absolute value of ΔΔG relative to the oscillation amplitude,

ΔU. (B) The surface oscillation efficiency (ε, color map) of a square wave operating at 1.0 Hz with varying amplitude

and oscillation amplitude endpoint exhibits two bands of increased efficiency (α ~ 0.6, β ~ 102 kJ mol-1, γ ~ 0.5, δ ~

1.4 kJ mol-1, and ΔHovr ~ 0 kJ mol-1).

Page 20: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 19

controlling the mobility of metal ions across

membranes against equilibrium[74]. These

molecular machines and pumps are relevant to the

dynamic oscillating catalyst surface concept

addressed in this work due to a common

mechanism. Of the many methods of manipulating

molecules, many molecular machines and pumps

utilize energy to implement a dynamic ‘pulsating’

ratchet energy profile[75,76,77]. Similar to Figure 14,

progression along a pulsating ratchet permits

molecules to spontaneously access the current low

energy states; the input of energy into the ratchet

system alters the free energy landscape permitting

molecules to access new low energy states that

constitute progress in position (i.e., molecular

pumps). The evolution of changing energy profiles

biases progress in one direction based on local peak

heights, consistent with its ‘ratchet’ name[73].

Molecular pumps and ‘catalytic pumps’ share a

common purpose; move molecules against

thermodynamics. Molecular pumps move

molecules against the free energy gradient

associated with concentration or pressure, while the

catalytic pumps described here aim to advance the

extent of reaction against the chemical potential

associated with chemical reaction equilibrium.

Molecular pumps utilize the ratchet mechanism,

while the catalytic pump utilizes the ratchet

mechanism plus the addition of a surface reaction

as one of the free energy profile transition states.

Similarly, molecular pumps utilize external energy

sources (e.g. ATP in biological membrane

transport), while catalytic pumps oscillate binding

energy by a variety of proposed methods as

described in Figure 4 (e.g., oscillating surface strain

or electric field).

4.0 Conclusions. The dynamic catalytic promotion

of A-to-B surface reactions permits order-of-

magnitude rate enhancement and control of extent

of conversion for a broad set of surface

mechanisms. Classification of surface mechanisms

and their associated kinetic behavior requires

detailed description of the key parameters defined

here as gamma (γ, the ratio of variation in surface

adsorbate binding energy) and delta (δ, the surface

adsorbate binding energy common to two

intermediates). High gamma catalytic systems (γ >

1) achieve surface resonance between an oscillating

external stimulus and an oscillating surface reaction

by selecting a catalyst and dynamic conditions that

maintain two oscillating states of A* and B*,

respectively. Alternatively, low gamma catalytic

systems (γ < 1) require one of the two oscillating

states to consist of a completely bare surface (open

sites only). The two different dynamic catalytic

behaviors are visually apparent in their Sabatier

volcano shapes; superposition of numerous

Sabatier volcanoes with varying γ comprise a

‘supervolcano’ where the regime of surface-

reaction rate control is visible as deviation from the

superVolcano border. Dynamic catalysis was also

show to serve as a ‘catalytic molecular pump’ by

altering the binding energy of surface adsorbate

species via external work. The extent of work

converted into catalytic conversion deviating from

equilibrium was defined as the dynamic catalytic

efficiency. For both low and high γ catalytic

systems, conditions of efficient catalytic promotion

as high as 30-40% were identified as a strong

function of the selected dynamic oscillation

conditions of frequency, amplitude, and waveform

type.

Acknowledgements. We acknowledge financial

support of the Catalysis Center for Energy

Innovation, a U.S. Department of Energy - Energy

Frontier Research Center under Grant DE-

SC0001004. The authors acknowledge the

Minnesota Supercomputing Institute (MSI) at the

University of Minnesota for providing resources

that contributed to the research results reported

within this paper. URL: http://www.msi.umn.edu/

We acknowledge helpful discussions with

Professors Dan Frisbie, Michael Tsapatsis, and

Dionisios Vlachos.

Keywords. Catalysis, Sabatier, Dynamics,

Frequency, Resonance, Volcano, Ammonia

Supporting Information. Additional information

including computer code, time-on-stream data, and

simulation methods are included in the supporting

information.

Page 21: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 20

References

1 U.S. Department of Energy. “Basic Research Needs

for Catalysis Science to Transform Energy

Technologies: Report from the U.S. Department of

Energy, Office of Basic Energy Sciences Workshop

May 8-10, 2017, in Gaithersburg, Maryland”.

Retreived Feb. 3, 2017. Link:

https://science.energy.gov/~/media/bes/pdf/reports/20

17/BRN_CatalysisScience_rpt.pdf 2 C.L. Williams, C.C. Chang, P. Do, N. Nikbin, S.

Caratzoulas, D.G. Vlachos, R.F. Lobo, W. Fan, P.J.

Dauenhauer, “Cycloaddition of biomass-derived

furans for catalytic production of renewable p-

xylene,” ACS Catalysis 2012, 2(6), 935-939. 3 O.A. Abdelrahman, D.S. Park, K.P. Vinter, C.S.

Spanjers, L. Ren, H.J. Cho, K. Zhang, W. Fan, M.

Tsapatsis, P.J. Dauenhauer, “Renewable isoprene by

sequential hydrogenation of itaconic acid and

dehydra-decyclization of 3-methyl-tetrahydrofuran,”

ACS Catalysis 2017, 7(2), 1428-1431. 4 D.J. Saxon, M. Nasiri, M. Mandal, S. Maduskar, P.J.

Dauenhauer, C.J. Cramer, A.M. LaPointe, T.M.

Reineke, “Architectural control of isosorbide-based

polyethers via ring-opening polymerization,” Journal

of the American Chemical Society 2019. DOI:

10.1021/jacs.9b00083 5 J.A. Rodriguez, P. Liu, D.J. Stacchiola, S.D.

Senanayake, M.G. White, J.G. Chen, “Hydrogenation

of CO2 to methanol: Importance of metal-oxide and

metal-carbide interfaces in the activation of CO2,”

ACS Catalysis 2015, 5(11), 6696-6706. 6 M. Dunwell, W. Luc, Y. Yan, F. Jiao, B. Xu,

“Understanding surface-mediated electrochemical

reactions: CO2 reduction and beyond,” ACS

Catalysis 2018, 8(9), 8121-8129. 7 M. Hattori, T. Mori, Y. Inoue, M. Sasase, T. Tada, M.

Kitano, T. Yokoyama, M. Hara, H. Hosono,

“Enhanced catalytic ammonia synthesis with

transformed BaO,” ACS Catalysis 2018, 8(12),

10977-10984. 8 B. Lin, L. Heng, B. Fang, H. Yin, J. Ni, X. Wang, J.

Lin, L. Jiang, “Ammonia synthesis activity of

alumina-supported ruthenium catalyst enhanced by

alumina phase transformation,” ACS Catalysis 2019,

9(3), 1635-1644. 9 A.R. Singh, J.H. Montoya, B.A. Rohr, C. Tsai, A.

Vojvodic, J.K. Norskov, “Computational design of

active site structures with improved transition-state

scaling for ammonia synthesis,” ACS Catalysis 2018,

8(5), 4017-4024. 10 J.W. Erisman, M.A. Sutton, J. Galloway, Z. Klimont,

W. Winiwarter, “How a century of ammonia

synthesis changed the world,” Nature Geoscience

2008, 1, 636-639.

11 U. Deka, I. Lezcano-Gonzalez, B.M. Weckhuysen,

A.M. Beale, “Local environment and nature of Cu

active sites in zeolite-based catalysts for the selective

catalytic reduction of NOx,” ACS Catalysis 2013,

3(3), 413-427. 12 H. Yuan, J. Chen, H. Wang, P. Hu, “Activity trend

for low-concentration NO oxidation at room

temperature on rutile-type metal oxides,” ACS

Catalysis 2018, 8(11), 10864-10870. 13 J. Liu, D. Zhu, Y. Zheng, A. Vasileff, S. Qjao, “Self-

supported earth-abundant nanoarrays as efficient and

robust electrocatalysts for energy-related reactions,”

ACS Catalysis 2018, 8(7), 6707-6732. 14 J.E. Sutton, D.G. Vlachos, “A theoretical and

computational analysis of linear free energy relations

for the estimation of activation energies,” ACS

Catalysis 2012, 2, 1624-1634. 15 R.R. Chianelli, G. Berhault, P. Raybaud, S.

Kasztelan, J. Hafner, H. Toulhoat, “Periodic trends in

hydrodesulfurization: in support of the Sabatier

principle,” Applied Catalysis A: General 2002,

227(1-2), 83-96. 16 A. Logadottir, T.H. Rod, J.K. Norskov, B. Hammer,

S. Dahl, C.J.H. Jacobsen, “The Bronsted-

EvansPolyanyi relation and the volcano plot for

ammonia synthesis over transition metal catalysts,”

Journal of Catalysis 2001, 197, 229-231. 17 S. Ichikawa, “Volcano-shaped curves in

heterogeneous catalysis,” Chemical Engineering

Science 1990, 45(2), 529-535. 18 A.K. Vijh, “Sabatier-Balandin interpretation of the

catalytic decomposition of nitrous oxide on

metaloxide semiconductors,” Journal of Catalysis

1973, 31, 51-54. 19 D.A. Hansgen, D.G. Vlachos, J.G. Chen, “Using first

principles to predict bimetallic catalysts for the

ammonia decomposition reaction,” Nature Catalysis

2010, 2(6), 484-489. 20 A.W. Ulissi, A.J. Medford, T. Bligaard, J.K.

Norskov, “To address surface reaction network

complexity using scaling relations maching learning

and DFT calculations,” Nature Communications

2017, 8, 14621. 21 J.E. Sutton, D.G. Vlachos, “Effect of errors in linear

scaling relations and Bronsted-Evans-Polanyi

relations on activity and selectivity maps,” Journal of

Catalysis 2016, 338, 273-283. 22 J.K. Norskov, T. Bligaard, J. Rossmeisl, C.H.

Christensen, “Towards the computational design of

solid catalysts,” Nature Chemistry 2009, 1, 37-46. 23 F. Calle-Vallejo, D. Loffreda, M.T.M. Koper, P.

Sautet, “Introducing structural sensitivity into

Page 22: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 21

adsorption-energy scaling relations,” Nature

Chemistry 2015, 7, 403-410. 24 W.A. Ardagh, O. Abdelrahman, P.J. Dauenhauer,

“Principles of Dynamic Heterogeneous Catalysis:

Surface Resonance and Turnover Frequency

Response,” ACS Catalysis 2019, 9, 6929-6937. 25 A.A. Balandin, V.A. Ferapontov, A.A.

Tolstopyatova, “The activity of cadmium oxide as a

catalyst for hydrogen dehydrogenation,” Bulletin of

the Academy of Sciences of the USSR, Division of

Chemical Science. 1960, 9(10), 1630-1636. 26 A.A. Balandin, “The multiplet theory of catalysis.

Energy factors in catalysis,” 1964, 33(5), 549-579. 27 J.R. Jennings. Catalytic Ammonia Synthesis:

Fundamentals and Practice. 1991, Springer Science,

New York. DOI: 10.1007/978-1-4757-9592-9.

ISBN 978-1-4757-9594-3 28 J.L. Colby, P.J. Dauenhauer, B.C. Michael, A. Bhan,

L.D. Schmidt, “Improved utilization of biomass-

derived carbon by co-processing with hydrogen-rich

feedstocks in millisecond reactors,” Green Chemistry

2010, 12, 378-380. 29 L.C. Grabow, M. Mavrikakis, “Mechanism of

methanol synthesis on Cu through CO2 and CO

hydrogenation,” ACS Catalysis 2011, 1(4), 365-384. 30 D. Pakhare, J. Spivey, “A review of dry (CO2)

reforming of methane over noble metal catalysts,”

Chemical Society Reviews 2014, 43, 7813-7837. 31 Q. Fu, H. Saltsburg, M. Flytzani-Stephanopoulos,

“Active nonmetallic Au and Pt species on Ceria-

based water-gas shift catalysts,” Science 2003,

301(5635), 935-938. 32 G. Marnellos, M. Stoukides, “Ammonia synthesis at

atmospheric pressure,” Science 1998, 282(5386), 98-

100. 33 G. Ertl, “Surface Science and Catalysis – Studies on

the Mechanism of Ammonia Synthesis: The P.H.

Emmett Award Address,” Catalysis Reviews Science

and Engineering 1980, 21(2), 201-223. 34 C. Ratnasamy, J.P. Wagner, “Water Gas Shift

Catalysis,” Catalysis Reviews 2009, 51, 325-440. 35 D.S. Newsome, “The Water-Gas Shift Reaction,”

Catalysis Reviews Science and Engineering 1980,

21(2), 275-318. 36 F. Cavani, N. Ballarini, A. Cericola, “Oxidative

dehydrogenation of ethane and propane: How far

from commercial implementation?” Catalysis Today

2007, 127(1-4), 113-131. 37 M. Malmali, Y. Wei, A. McCormick, E.L. Cussler,

“Ammonia synthesis at reduced pressure via reactive

separation,” Industrial & Engineering Chemistry

Research 2016, 55(33), 8922-8932. 38 C. Smith, A.V. McCormick, E.L. Cussler,

“Optimizing the conditions for ammonia production

using absorption,” ACS Sustainable Chemistry &

Engineering 2019, 7(4), 4019-4029.

39 C.L. Williams, C.C. Chang, P. Do, N. Nikbin, S.

Caratzoulas, D.G. Vlachos, R.F. Lobo, W. Fan, P.J.

Dauenhauer, “Cycloaddition of biomass-derived

furans for catalytic production of renewable p-

xylene,” ACS Catalysis 2012, 2(6), 935-939. 40 C.L. Williams, K.P. Vinter, C.C. Chang, R. Xiong,

S.K. Green, S.I. Sandler, D.G. Vlachos, W. Fan, P.J.

Dauenhauer, “Kinetic Regimes in the tandem

reactions of H-BEA catalyzed formation of p-xylene

from dimethylfuran,” Catalysis Science and

Technology 2016, 6, 178-187. 41 L.C. Buelens, V.V. Galvita, H. Poelman, C.

Detavernier, G.B. Marin, “Super-dry reforming of

methane intensifies CO2 utilization via Le Chatelier’s

principle,” Science 2016, 354(6311), 449-452. 42 W. Lin, C.J. Murphy, “A demonstration of Le

Chatelier’s Principle on the nanoscale,” ACS Central

Science 2017, 3(10), 1096-1102. 43 I. Roger, M.A. Shipman, M.D. Symes, “Earth-

abundant catalysts for electrochemical and

photoelectrochemical water splitting,” Nature

Reviews Chemistry 2017, 1, 0003. 44 D. Todd, M. Schwager, W. Merida,

“Thermodynamics of high-temperature, high-

pressure water electrolysis,” Journal of Power

Sources 2014, 269, 424-429. 45 B.H.R. Suryanto, H.L. Du, D. Wang, J. Chen, A.N.

Simonov, D.R. MacFarlane, “Challenges and

prospects in the catalysis of electroreduction of

nitrogen to ammonia,” Nature Catalysis 2019, 2,

290-296. 46 A. Ardagh, O. Abdelrahman, P.J. Dauenhauer,

“Principles of Dynamic Heterogeneous Catalysis:

Surface Resonance and Turnover Frequency

Response,” ChemRxiv, March 1, 2019.

doi.org/10.26434/chemrxiv.7790009.v1 47 J.E. Sutton, D.G. Vlachos, “A theoretical and

computational analysis of linear free energy relations

for the estimation of activation energies,” ACS

Catalysis 2012, 2(8), 1624-1634. 48 M. Mavrikakis, B. Hammer, J.K. Norskov, “Effect of

strain on the reactivity of metal surfaces,” Physical

Review Letters 1998, 81(13), 2819. 49 Y. Xu, M. Mavrikakis, “Adsorption and dissociation

of O2 on Cu(111): thermochemistry, reaction barrier

and the effect of strain,” Surface Science 2001, 494,

131-144. 50 Z. Xia, S. Guo, “Strain engineering of metal-based

nanomaterials for energy electrocatalysis,” Chemical

Society Reviews, 2019, 48, 3265-3278. DOI:

10.1039/c8cs00846a 51 J. Weissmuller, “Adsorption-strain coupling at solid

surfaces,” Current Opinion in Chemical Engineering

2019, 24, 45-53. 52 A.J. Bennett, “The effect of applied electric fields on

chemisorption,” Surface Science 1975, 50, 77-94.

Page 23: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

____________________________________________________________________________ Ardagh, et al. Page 22

53 M.T.M. Koper, R.A. van Santen, “Electric field

effects on CO and NO adsorption at the Pt(111)

surface,” Journal of Electroanalytical Chemistry

1999, 476, 64-70. 54 S.A. Wasileski, M.T.M. Koper, M.J. Weaver, “Field-

Dependent Chemisorption of Carbon Monoxide on

Platinum-Group (111) Surfaces: Relationships

between Binding Energetics, Geometries, and

Vibrational Properties as Assessed by Density

Functional Theory” J. Phys. Chem. B 2001, 105,

3518-3530. 55 Y. Wang, C.H. Kim, Y. Yoo, J.E. Johns, C.D.

Frisbie, “Field effect modulation of heterogeneous

charge transfer kinetics at back-gated twodimensional

MoS2 electrodes,” Nano Letters 2017, 17, 7586-

7592. 56 C.H. Kim, C.D. Frisbie, “Field effect modulation of

outer-sphere electrochemistry at back-gated, ultrathin

ZnO electrodes,” Journal of the American Chemical

Society 2016, 138, 7220-7223. 57 C.H. Kim, Y. Wang, C.D. Frisbie, “Continuous and

reversible tuning of electrochemical reaction kinetics

on back-gated 2D semiconductor electrodes: steady

state analysis using a hydrodynamic method,”

Analytical Chemistry 2019, 91, 1627-1635. 58 P. Deshlahra, W.F. Schneider, G.H. Bernstein, E.E.

Wolf, “Direct control of electron transfer to the

surface-CO bond on a Pt/TiO2 catalytic diode,”

Journal of the American Chemical Society 2011, 133,

16459-16467. 59 K. Hahn, “Atomic and molecular adsorption on

Re(0001),” Topics in Catalysis 2014, 57, 54-68. 60 J.A. Herron, S. Tonelli, M. Mavrikakis, “Atomic and

molecular adsorption on Ru(0001),” Surface Science

2013, 614, 64-74 61 D.C. Ford, Y. Xu, M. Mavrikakis, “Atomic and

molecular adsorption on Pt(111),” Surface Science

2005, 587, 159-174. 62 J.A. Herron, S. Tonelli, M. Mavrikakis, “Atomic and

molecular adsorption on Pd(111),” Surface Science

2012, 606, 1670-1679. 63 Y. Santiago-Rodriguez, J.A. Herron, M.C. Curet-

Arana, M. Mavrikakis, “Atomic and molecular

adsorption on Au(111),” Surface Science 2014, 627,

57-69. 64 F. Che, S. Ha, J.-S. McEwen, “Elucidating the field

influence on the energetics of the methane steam

reforming reaction: A density functional theory

study,” Applied Catalysis B: Environmental 2016,

195, 7-89. 65 F. Liu, T. Xue, C. Wu, S. Yang, “Coadsorption of

CO and O over strained metal surfaces,” Chemical

Physics Letters 2019, 722, 18-25. 66 Z.W. Ulissi, A.J. Medford, T. Bligaard, J.K.

Norskov, “To address surface reaction network

complexity using scaling relations machine learning

and DFT calculations,” Nature Communications

2017, 8, 14621. 67 E.R. Kay, D.A. Leigh, F. Zerbetto, “Synthetic

molecular motors and mechanical machines,”

Angewandte Chemie International Edition 2007, 46,

72-191. 68 S. Erbas-Cakmak, D.A. Leight, C.T. McTernan, A.L.

Nussbaumer, “Artificial Molecular Machines,”

Chemical Reviews 2015, 115, 10081-10206. 69 L. Zhang, V. Marcos, D.A. Leight, “Molecular

machines with bio-inspired mechanisms,” PNAS

2018, 115(38), 9397. 70 V. Blanco, D.A. Leigh, V. Marcos, “Artificial

switchable catalysts,” Chemical Society Reviews

2015, 44, 5341-5370. 71 J. Clayden, “No turning back for motorized

molecules,” Nature 2016, 534, 187. 72 M.R. Wilson, J. Sola, A. Carlone, S.M. Goldup, N.

Lebrasseur, D.A. Leigh, “An autonomous chemically

fueled small-molecule motor,” Nature 2016, 534,

235. 73 C. Cheng, P.R. McGonigal, S.T. Schneebeli, H. Li,

N.A. Vermeulen, C. Ke, J.F. Stoddart, “An artificial

molecular pump,” Nature Nanotechnology, 2015, 10,

547. 74 I.M. Bennett, H.M. Vanegas Farfano, F. Bogani, A.

Primak, P.A. Liddell, L. Otero, L. Sereno, J.J. Silber,

A.L. Moore, T.A. Moore, D. Gust, “Active transport

of Ca2+ by an artificial photosynthetic membrane”.

Nature 2002, 420, 398–401. 75 E.R. Kay, D.A. Leigh, “Rise of the Molecular

Machines,” Ang. Chem. Int. Ed. 2015, 54, 10080-

10088. 76 V. Serreli, C.F. Lee, E.R. Kay, D.A. Leigh, “A

molecular information ratchet,” Nature 2007, 445,

523-527. 77 R. D. Astumian, I. Derenyi, “Fluctuation driven

transport and models of molecular motors and

pumps,” Eur. Biophys. J. 1998, 27, 474-489.

Page 25: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S1

Electronic Supplementary Information

Catalytic Resonance Theory:

SuperVolcanoes, Catalytic Molecular Pumps, and

Oscillatory Steady State

M. Alexander Ardagh1,2, Turan Birol1, Qi Zhang1,

Omar A. Abdelrahman3, Paul J. Dauenhauer1,2*

1

University of Minnesota, Department of Chemical Engineering and Materials Science, 421 Washington

Ave. SE, Minneapolis, MN 55455.

2Catalysis Center for Energy Innovation, University of Delaware, 150 Academy Street, Colburn

Laboratory, Newark, DE 19716.

3University of Massachusetts Amherst, Department of Chemical Engineering, 686 N. Pleasant Street,

Amherst, MA 01003.

*Corresponding Author: [email protected]

# of Figures: 2

# of Tables: 8

# of Equations: 6

Table of Contents:

Section S1. Matlab 2017b and 2019a Code

● CSTR Model

● Batch Reactor Model

● Dynamic Catalysis CSTR Shell Code with Square Waveform

● Dynamic Catalysis Batch Reactor Shell Code with Square or Triangle Waveform

● Dynamic Catalysis Batch Reactor Shell Code with Sinusoidal Waveform

Section S2. Matlab ODE Solver Performance

● Static Catalysis

● Dynamic Catalysis

● Solvers:

○ ODE45

○ ODE15s

○ ODE23s

○ ODE23t

○ ODE23tb

○ Radau

Section S3. Static Catalysis Time on Stream Data

● CSTR at 1 % Yield of B

● Batch reactor at 50 % Yield of B

Section S4. Data from Heatmap Figures

Page 26: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S2

Section S5. Binding Energy Derivation

Section S1. Matlab 2017b and 2019a Code

CSTR Model

% CSTR model

% Description:

% Continuously Stirred Tank Reactor (CSTR) with reaction A -> B.

% The binding energies of A and B are the controls.

function xdot = cstr1(t,x)

global Rg R TK Caf Cbf qdot w u delHovr alpha beta gamma delta BEb0 delBEb

% Input (1):

% Number of catalytic active sites (gmol)

Ns = u;

% States (2):

% Concentration of A in CSTR (M)

Ca = x(1,1);

% Concentration of B in CSTR (M)

Cb = x(2,1);

% Amount of adsorbed A in CSTR (gmol)

Ca_star = x(3,1);

% Amount of adsorbed B in CSTR (gmol)

Cb_star = x(4,1);

% Number of vacant sites in CSTR (gmol)

C_star = Ns - Ca_star - Cb_star;

% Convert units

% Pressure of A in CSTR (bar)

Pa = Ca*Rg*TK;

% Pressure of B in CSTR (bar)

Pb = Cb*Rg*TK;

% Surface coverage of A in CSTR (unitless)

Theta_A_star = Ca_star/Ns;

% Surface coverage of B in CSTR (unitless)

Theta_B_star = Cb_star/Ns;

% Surface coverage of * in CSTR (unitless)

Theta_star = C_star/Ns;

% Reactor parameters:

% Volume of CSTR

% Catalyst density (g/mL)

rho = 3.58;

% Bed void fraction (unitless)

epsilon = 0.375;

% Volume calculation (L)

V = (w/1000)*(1/rho)*(1/1000)*(1/(1-epsilon));

Page 27: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S3

% Reaction parameters:

% Binding energies

% Volcano x-axis zero point (eV)

BEa0 = (BEb0 - ((1-gamma)*delta) - (delHovr/96.485e3))/gamma;

% Binding energy of B (J/gmol)

BEb = (BEb0 + delBEb)*96.485e3;

% Restrict binding energy of B to positive values

if BEb < 0

BEb = 0;

End

% Binding energy of A (J/gmol)

BEa = (BEa0 + ((1/gamma)*delBEb))*96.485e3;

% Restrict binding energy of A to positive values

if BEa < 0

BEa = 0;

End

% Heats of reaction (J/gmol)

delH1 = -BEa; % A -> A*

delH2 = delHovr + BEa - BEb; % A* -> B*

delH3 = BEb; % B* -> B

% Ea - Activation energies in the Arrhenius equation (J/gmol)

Ea1f = 0e3; % A -> A*

Ea2f = alpha*delH2 + beta; % A* -> B*

% Restrict activation energy to positive values

if Ea2f < 0

Ea2f = 0;

end

Ea3f = delH3; % B* -> B

% Pre-exponential factors

A1f = 1e6; % 1/bar-sec

A2f = 1e13; % 1/sec

A3f = 1e13; % 1/sec

% Equilibrium constants

K1 = 1e-7*exp(-delH1/(R*TK)); % 1/bar

K2 = 1*exp(-delH2/(R*TK)); % unitless

K3 = 1e7*exp(-delH3/(R*TK)); % bar

% Rate constants

k1f = A1f*exp(-Ea1f/(R*TK)); % 1/bar-sec

k1r = k1f/K1; % 1/sec

k2f = A2f*exp(-Ea2f/(R*TK)); % 1/sec

k2r = k2f/K2; % 1/sec

k3f = A3f*exp(-Ea3f/(R*TK)); % 1/sec

k3r = k3f/K3; % 1/bar-sec

Page 28: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S4

% Compute xdot:

xdot(1,1) = (qdot/V*(Caf - Ca) - k1f*Pa*Theta_star*(Ns/V) + k1r*Theta_A_star*(Ns/V)); % M/sec

xdot(2,1) = (qdot/V*(Cbf - Cb) + k3f*Theta_B_star*(Ns/V) - k3r*Pb*Theta_star*(Ns/V)); % M/sec

xdot(3,1) = k1f*Pa*Theta_star*Ns - k1r*Theta_A_star*Ns - k2f*Theta_A_star*Ns +

k2r*Theta_B_star*Ns; % gmol/sec

xdot(4,1) = - k3f*Theta_B_star*Ns + k3r*Pb*Theta_star*Ns + k2f*Theta_A_star*Ns -

k2r*Theta_B_star*Ns; % gmol/sec

Batch Reactor Model

% Batch reactor model

% Description:

% Batch reactor with reaction A -> B.

% The binding energies of A and B are the controls.

function xdot = batch1(t,x)

global Rg R TK w u delHovr alpha beta gamma delta BEb0 delBEb

% Input (1):

% Number of catalytic active sites (gmol)

Ns = u;

% States (2):

% Concentration of A in batch reactor (M)

Ca = x(1,1);

% Concentration of B in batch reactor (M)

Cb = x(2,1);

% Amount of adsorbed A in batch reactor (gmol)

Ca_star = x(3,1);

% Amount of adsorbed B in batch reactor (gmol)

Cb_star = x(4,1);

% Number of vacant sites in batch reactor (gmol)

C_star = Ns - Ca_star - Cb_star;

% Convert units

% Pressure of A in batch reactor (bar)

Pa = Ca*Rg*TK;

% Pressure of B in batch reactor (bar)

Pb = Cb*Rg*TK;

% Surface coverage of A in batch reactor (unitless)

Theta_A_star = Ca_star/Ns;

% Surface coverage of B in batch reactor (unitless)

Theta_B_star = Cb_star/Ns;

% Surface coverage of * in batch reactor (unitless)

Theta_star = C_star/Ns;

% Reactor parameters:

% Volume of batch reactor

% Catalyst density (g/mL)

rho = 3.58;

Page 29: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S5

% Bed void fraction (unitless)

epsilon = 0.375;

% Volume calculation (L)

V = (w/1000)*(1/rho)*(1/1000)*(1/(1-epsilon));

% Reaction parameters:

% Binding energies

% Volcano x-axis zero point (eV)

BEa0 = (BEb0 - ((1-gamma)*delta) - (delHovr/96.485e3))/gamma;

% Binding energy of B (J/gmol)

BEb = (BEb0 + delBEb)*96.485e3;

% Restrict binding energy of B to positive values

if BEb < 0

BEb = 0;

End

% Binding energy of A (J/gmol)

BEa = (BEa0 + ((1/gamma)*delBEb))*96.485e3;

% Restrict binding energy of A to positive values

if BEa < 0

BEa = 0;

End

% Heats of reaction (J/gmol)

delH1 = -BEa; % A -> A*

delH2 = delHovr + BEa - BEb; % A* -> B*

delH3 = BEb; % B* -> B

% Ea - Activation energies in the Arrhenius equation (J/gmol)

Ea1f = 0e3; % A -> A*

Ea2f = alpha*delH2 + beta; % A* -> B*

% Restrict activation energy to positive values

if Ea2f < 0

Ea2f = 0;

end

Ea3f = delH3; % B* -> B

% Pre-exponential factors

A1f = 1e6; % 1/bar-sec

A2f = 1e13; % 1/sec

A3f = 1e13; % 1/sec

% Equilibrium constants

K1 = 1e-7*exp(-delH1/(R*TK)); % 1/bar

K2 = 1*exp(-delH2/(R*TK)); % unitless

K3 = 1e7*exp(-delH3/(R*TK)); % bar

% Rate constants

k1f = A1f*exp(-Ea1f/(R*TK)); % 1/bar-sec

k1r = k1f/K1; % 1/sec

Page 30: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S6

k2f = A2f*exp(-Ea2f/(R*TK)); % 1/sec

k2r = k2f/K2; % 1/sec

k3f = A3f*exp(-Ea3f/(R*TK)); % 1/sec

k3r = k3f/K3; % 1/bar-sec

% Compute xdot:

xdot(1,1) = -k1f*Pa*Theta_star*(Ns/V) + k1r*Theta_A_star*(Ns/V); % M/sec

xdot(2,1) = k3f*Theta_B_star*(Ns/V) - k3r*Pb*Theta_star*(Ns/V); % M/sec

xdot(3,1) = k1f*Pa*Theta_star*Ns - k1r*Theta_A_star*Ns - k2f*Theta_A_star*Ns +

k2r*Theta_B_star*Ns; % gmol/sec

xdot(4,1) = - k3f*Theta_B_star*Ns + k3r*Pb*Theta_star*Ns + k2f*Theta_A_star*Ns -

k2r*Theta_B_star*Ns; % gmol/sec

Dynamic Catalysis CSTR Shell Code with Square Waveform

% Remove prior data and runs

clear

clc

% Step test for Model 1 - CSTR

global Rg R TK Caf Cbf qdot w u delHovr alpha beta gamma delta BEb0 delBEb

% Constants:

% Gas constant

Rg = 8.314459848e-2; % L-bar/K-gmol

R = 8.314459848; % J/K-gmol

% Reactor parameters:

% Temperature

Tc = 150; % deg C

TK = Tc + 273.15; % K

% Feed pressure (bar)

Paf = 100;

Pbf = 0;

% Feed concentration (M)

Caf = Paf/(Rg*TK);

Cbf = Pbf/(Rg*TK);

% Volumetric flowrate

qsdot = 50; % mL/min

qdot = qsdot/60000; % L/sec

% Steady state initial conditions for the states

Ca_ss = Caf; % M

Cb_ss = Cbf; % M

Ca_star_ss = 0; % gmol

Cb_star_ss = 0; % gmol

x_ss = [Ca_ss;Cb_ss;Ca_star_ss;Cb_star_ss];

Page 31: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S7

% Steady state initial condition for the control

u_ss = 0; % gmol active sites

% Open loop step change

% Catalyst weight (mg)

w = 138;

% Active site density (gmol/g catalyst)

srho = 20e-6;

% Active site calculation (gmol)

u = (w/1000)*srho;

% Reaction parameters:

% Overall heat of reaction for A -> B (J/gmol)

delHovr = 0e3;

% Brønsted-Evans-Polanyi relationship parameters

alpha = 0.6; % unitless

beta = 102e3; % J/gmol

% Binding energy relationship parameters

gamma = 0.5; % unitless

delta = 1.4; % eV

% Volcano plot parameter:

% x-axis zero point (eV)

BEb0 = delta + delHovr/96.485e3;

% Volcano plot peak calculation

% Iterate binding energies of A and B (eV)

is = -delta;

ii = 0.01;

ie = delta;

% Preallocate BEb, TOF, and surface coverage matrices

je = ((ie - is)/ii + 1); % final index

delBEbv = zeros(round(je),1); % eV

TOFbvf = zeros(round(je),1); % 1/sec

Theta_A_starvf = zeros(round(je),1); % unitless

Theta_B_starvf = zeros(round(je),1); % unitless

% Begin iteration (eV)

for i = is:ii:ie

% Set binding energy of B (eV)

delBEb = i; % eV

% Iterate at each binding energy

for n = 1:inf % index

% Final Time (sec)

tfv = 5e100;

% Generate CSTR data

opts = odeset('RelTol’,1e-8,’AbsTol’,1e-9);

[tv,xv] = ode15s('cstr1',[0 tfv],x_ss,opts); % Time and state outputs

% Parse out the state values (M)

Page 32: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S8

Cav = xv(:,1);

Cbv = xv(:,2);

% Measure reactor performance

Ybv = (Cbv./(Cav + Cbv))*100; % mol %

% Converge on C yield of B (mol %)

C = 1;

if abs(Ybv(end) - C) > 0.01 % Convergence criterion

qdot = qdot*Ybv(end)/C; % L/sec

nv = n; % Indicates that volcano criterion was not met

nv; % Display iteration count

clear tv xv Cav Cbv Ybv

else

% Compute matrix index

j = ((i - is)/ii) + 1; % unitless

% Parse out state values (gmol)

Ca_starv = xv(:,3);

Cb_starv = xv(:,4);

% Measure reactor performance

TOFbv = Cbv*qdot/u; % 1/sec

% Convert units

Theta_A_starv = Ca_starv/u; % unitless

Theta_B_starv = Cb_starv/u; % unitless

% Store volcano plot results

delBEbv(round(j)) = i; % eV

TOFbvf(round(j)) = TOFbv(end); % 1/sec

Theta_A_starvf(round(j)) = Theta_A_starv(end); % unitless

Theta_B_starvf(round(j)) = Theta_B_starv(end); % unitless

clear tv xv Cav Cbv Ybv Ca_starv Cb_starv TOFbv Theta_A_starv Theta_B_starv

break

end

end

end

% Extract volcano plot peak information

peakTOF = max(TOFbvf); % 1/sec

% Preallocate relative binding energy matrix (eV)

peakdelBEb = zeros(round(je),1);

% Iterate relative binding energy of B (eV)

for k = 1:je % index

if abs(TOFbvf(k) - peakTOF) < peakTOF/100 % Convergence criterion

peakdelBEb(k) = delBEbv(k); % Store relative binding energy

end

end

Page 33: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S9

% Calculate volcano plot peak position (eV)

peakdelBEb = mean(nonzeros(peakdelBEb));

% Display volcano plot peak results

peakdelBEb; % eV

peakTOF; % 1/sec

% Transition to dynamic catalysis

% Dynamic catalysis parameters

% Oscillation frequency

tau2 = 5e-4; % sec

taur = 1.0; % unitless

tau1 = taur*tau2; % sec

fosc = 1/(tau1 + tau2); % Hz

% Number of oscillations (unitless)

Nosc = round(fosc) + 1;

% Set minimum number of oscillations

if Nosc < 11

Nosc = 11;

end

% Oscillation endpoints (eV)

delBEb0 = peakdelBEb;

delU = 0.6;

UR = delBEb0 + delU/2.0;

UL = UR - delU;

% Static catalysis calculation

% Starting time

ts = 0;

% Final time

te = 5e3;

% Generate CSTR data

% Reset volumetric flowrate (L/sec)

qdot = qsdot/60000;

% Static catalysis at volcano plot peak

delBEb = delBEb0; % eV

% Start static catalysis timer (sec)

tic

% Iterate until convergence

for n = 1:inf % index

% Generate time span matrix (sec)

% Time interval

ti = te/40;

% Time span matrix

tspan = ts:ti:te;

[t,x] = ode15s('cstr1',tspan,x_ss,opts); % Time and state outputs

Page 34: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S10

% Parse out the state values

Ca(:,1) = x(:,1); % M

Cb(:,1) = x(:,2); % M

% Measure reactor performance

Yb = (Cb./(Ca + Cb))*100; % mol %

% Converge on C yield of B (mol %)

if abs(Yb(end) - C) > 0.01 % Convergence criterion

qdot = qdot*Yb(end)/C; % L/sec

ns = n; % Indicates that static catalysis criterion was not met

ns; % Display iteration count

clear t x Ca Cb Yb

else

% Stop static catalysis timer (sec)

toc

% Store static catalysis results

tsv(:,1) = t; % sec

Ca_star(:,1) = x(:,3); % gmol

Cb_star(:,1) = x(:,4); % gmol

TOFb = Cb*qdot/u; % 1/sec

% Remove static catalysis data and run

clear t x

% Reset volumetric flowrate (L/sec)

qdot = qsdot/60000;

break

end

end

% Begin dynamic catalysis

% Initial conditions for dynamic catalysis

x_ss(:,2) = [Ca(end,1);Cb(end,1);Ca_star(end,1);Cb_star(end,1)];

% Start dynamic catalysis timer (sec)

tic

% Iterate until convergence

for n = 1:inf % index

% Begin dynamic catalysis

for l = 2:Nosc % index

% Even numbered oscillations

if mod(l,2) == 0

% Operate at strong binding (eV)

delBEb = UR;

Page 35: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S11

% Construct new time span matrix (sec)

% Final time

te = tau1;

% Time interval

ti = te/40;

% Time span matrix

tspan = ts:ti:te;

% Generate CSTR data

[t,x] = ode15s('cstr1',tspan,x_ss(:,l),opts); % Time and state outputs

% Parse out the state values

tsv(:,l) = t + tsv(end,l-1); % sec

Ca(:,l) = x(:,1); % M

Cb(:,l) = x(:,2); % M

Ca_star(:,l) = x(:,3); % gmol

Cb_star(:,l) = x(:,4); % gmol

% Initial conditions for the next oscillation

x_ss(:,l+1) = [Ca(end,l);Cb(end,l);Ca_star(end,l);Cb_star(end,l)];

clear t x

% Odd numbered oscillations

else

% Operate at weak binding (eV)

delBEb = UL;

% Construct new time span matrix (sec)

% Final time

te = tau2;

% Time interval

ti = te/40;

% Time span matrix

tspan = ts:ti:te;

% Generate CSTR data

[t,x] = ode15s('cstr1',tspan,x_ss(:,l),opts); % Time and state outputs

% Parse out the state values

tsv(:,l) = t + tsv(end,l-1); % sec

Ca(:,l) = x(:,1); % M

Cb(:,l) = x(:,2); % M

Ca_star(:,l) = x(:,3); % gmol

Cb_star(:,l) = x(:,4); % gmol

% Initial conditions for the next oscillation

x_ss(:,l+1) = [Ca(end,l);Cb(end,l);Ca_star(end,l);Cb_star(end,l)];

clear t x

end

end

% End dynamic catalysis

% Calculate state values (gmol)

Page 36: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S12

C_star = u - Ca_star - Cb_star;

% Convert units

Pa = Ca*Rg*TK; % bar

Pb = Cb*Rg*TK; % bar

Theta_A_star = Ca_star/u; % unitless

Theta_B_star = Cb_star/u; % unitless

Theta_star = C_star/u; % unitless

% Measure reactor performance

Yb = (Cb./(Ca + Cb))*100; % mol %

% Preallocate TOF matrix

TOFb = [TOFb(:,1),zeros((te/ti)+1,Nosc-1)]; % 1/sec

TOFb(:,2:end) = Cb(:,2:end)*qdot/u; % 1/sec

% Converge on C yield of B (mol %)

% Data at the end of the run

% Time data (sec)

tdiff = [tsv(:,end-1);tsv(:,end)];

% Yield data (mol %)

Ybdiff = [Yb(:,end-1);Yb(:,end)];

% Data in the middle of the run

% Time data (sec)

tdiff2 = [tsv(:,(end-1)/2);tsv(:,(end-1)/2+1)];

% Yield data (mol %)

Ybdiff2 = [Yb(:,(end-1)/2);Yb(:,(end-1)/2+1)];

% Integrate yield of B over 1 oscillation period

% Preallocate integral matrix (sec*mol %)

Ybint = zeros(size(Ybdiff));

% Integration with midpoint Riemann sum

for m = 2:size(tdiff,1)

Ybint(m) = (tdiff(m) - tdiff(m-1))*mean([Ybdiff(m-1);Ybdiff(m)]);

End

% Preallocate integral matrix (sec*mol %)

Ybint2 = zeros(size(Ybdiff2));

% Integration with midpoint Riemann sum

for p = 2:size(tdiff2,1)

Ybint2(p) = (tdiff2(p) - tdiff2(p-1))*mean([Ybdiff2(p-1);Ybdiff2(p)]);

end

% Calculate time-averaged yield of B (mol %)

Ybavg = sum(Ybint)*fosc;

Ybavg2 = sum(Ybint2)*fosc;

% Check for convergence

if abs(Ybavg - Ybavg2) > min([Ybavg,Ybavg2])/100 % Convergence criterion 1

Nosc = round(2*Nosc) + 1; % Adjust number of oscillations

Page 37: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S13

nn = n; % Indicates that criterion 1 was not met

nn; % Display iteration number

clear tsv(:,2:end) Ca(:,2:end) Cb(:,2:end) Ca_star(:,2:end) Cb_star(:,2:end) x_ss(:,3:end)

else

if abs(Ybavg - C) > 0.01 % Convergence criterion 2

qdot = qdot*Ybavg/C; % Adjust volumetric flowrate

nq = n; % Indicates that criterion 2 was not met

nq; % Display iteration number

clear tsv(:,2:end) Ca(:,2:end) Cb(:,2:end) Ca_star(:,2:end) Cb_star(:,2:end) x_ss(:,3:end)

else

% Stop dynamic catalysis timer (sec)

toc

break

end

end

end

% Integrate TOF over 1 oscillation period

TOFbdiff = [TOFb(:,end-1);TOFb(:,end)];

% Preallocate TOF integral matrix (unitless)

TOFbint = zeros(size(TOFbdiff));

% Integration with midpoint Riemann sum

for m = 2:size(tdiff,1)

TOFbint(m) = (tdiff(m) - tdiff(m-1))*mean([TOFbdiff(m-1);TOFbdiff(m)]);

end

% Calculate the time-averaged TOF (1/sec)

TOFbavg = sum(TOFbint)*fosc;

% Display final results

plot(tsv,Yb) % Visually check for convergence

TOFbavg; % 1/sec

Dynamic Catalysis Batch Reactor Shell Code with Square or Triangle Waveform

% Remove prior data and runs

clear

clc

% Step test for models - CSTR and Batch Reactor

global Rg R TK Caf Cbf qdot w u delHovr alpha beta gamma delta BEb0 delBEb

% Constants:

% Gas constant

Rg = 8.314459848e-2; % L-bar/K-gmol

R = 8.314459848; % J/K-gmol

% Reactor parameters:

% Temperature

Tc = 150; % deg C

Page 38: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S14

TK = Tc + 273.15; % K

% Feed pressure (bar)

Paf = 100;

Pbf = 0;

% Feed concentration (M)

Caf = Paf/(Rg*TK);

Cbf = Pbf/(Rg*TK);

% Volumetric flowrate

qsdot = 50; % mL/min

qdot = qsdot/60000; % L/sec

% Steady state initial conditions for the states

Ca_ss = Caf; % M

Cb_ss = Cbf; % M

Ca_star_ss = 0; % gmol

Cb_star_ss = 0; % gmol

x_ss = [Ca_ss;Cb_ss;Ca_star_ss;Cb_star_ss];

% Steady state initial condition for the control

u_ss = 0; % gmol active sites

% Open loop step change

% Catalyst weight (mg)

w = 138;

% Active site density (gmol/g catalyst)

srho = 20e-6;

% Active site calculation (gmol)

u = (w/1000)*srho;

% Reaction parameters:

% Overall heat of reaction for A -> B (J/gmol)

delHovr = 0e3;

% Brønsted-Evans-Polanyi relationship parameters

alpha = 0.6; % unitless

beta = 102e3; % J/gmol

% Binding energy relationship parameters

gamma = 0.5; % unitless

delta = 1.4; % eV

% Volcano plot parameter:

% x-axis zero point (eV)

BEb0 = delta + delHovr/96.485e3;

% Volcano plot peak calculation

% Iterate binding energies of A and B (eV)

is = -delta;

ii = 0.01;

ie = delta;

% Preallocate BEb, TOF, and surface coverage matrices

je = ((ie - is)/ii + 1); % final index

Page 39: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S15

delBEbv = zeros(round(je),1); % eV

TOFbvf = zeros(round(je),1); % 1/sec

Theta_A_starvf = zeros(round(je),1); % unitless

Theta_B_starvf = zeros(round(je),1); % unitless

% Begin iteration (eV)

for i = is:ii:ie

% Set binding energy of B (eV)

delBEb = i; % eV

% Iterate at each binding energy

for n = 1:inf % index

% Final Time (sec)

tfv = 5e100;

% Generate CSTR data

opts = odeset('RelTol',1e-8,'AbsTol',1e-9);

[tv,xv] = ode15s('cstr1',[0 tfv],x_ss,opts); % Time and state outputs

% Parse out the state values (M)

Cav = xv(:,1);

Cbv = xv(:,2);

% Measure reactor performance

Ybv = (Cbv./(Cav + Cbv))*100; % mol %

% Converge on C yield of B (mol %)

C = 1;

if abs(Ybv(end) - C) > 0.01 % Convergence criterion

qdot = qdot*Ybv(end)/C; % L/sec

nv = n; % Indicates that volcano criterion was not met

nv; % Display iteration count

clear tv xv Cav Cbv Ybv

else

% Compute matrix index

j = ((i - is)/ii) + 1; % unitless

% Parse out state values (gmol)

Ca_starv = xv(:,3);

Cb_starv = xv(:,4);

% Measure reactor performance

TOFbv = Cbv*qdot/u; % 1/sec

% Convert units

Theta_A_starv = Ca_starv/u; % unitless

Theta_B_starv = Cb_starv/u; % unitless

% Store volcano plot results

delBEbv(round(j)) = i; % eV

TOFbvf(round(j)) = TOFbv(end); % 1/sec

Theta_A_starvf(round(j)) = Theta_A_starv(end); % unitless

Page 40: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S16

Theta_B_starvf(round(j)) = Theta_B_starv(end); % unitless

clear tv xv Cav Cbv Ybv Ca_starv Cb_starv TOFbv Theta_A_starv Theta_B_starv

break

end

end

end

% Extract volcano plot peak information

peakTOF = max(TOFbvf); % 1/sec

% Preallocate relative binding energy matrix (eV)

peakdelBEb = zeros(round(je),1);

% Iterate relative binding energy of B (eV)

for k = 1:je % index

if abs(TOFbvf(k) - peakTOF) < peakTOF/100 % Convergence criterion

peakdelBEb(k) = delBEbv(k); % Store relative binding energy

end

end

% Calculate volcano plot peak position (eV)

peakdelBEb = mean(nonzeros(peakdelBEb));

% Display volcano plot peak results

peakdelBEb; % eV

peakTOF; % 1/sec

% Transition to dynamic catalysis

% Dynamic catalysis parameters

% Oscillation frequency

tau2 = 5e0; % sec

taur = 1.0; % unitless

tau1 = taur*tau2; % sec

fosc = 1/(tau1 + tau2); % Hz

% Number of oscillations (unitless)

Nosc = round(fosc) + 1;

% Set minimum number of oscillations

if Nosc < 11

Nosc = 11;

end

% Oscillation endpoints (eV)

delBEb0 = peakdelBEb;

delU = 1.50;

UR = delBEb0 + delU/2.0;

UL = UR - delU;

% Static catalysis calculation

% Starting time (sec)

ts = 0;

% Final time (sec)

te = 5e3;

Page 41: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S17

% Generate batch reactor data

% Static catalysis at volcano plot peak

delBEb = delBEb0; % eV

% Start static catalysis timer (sec)

tic

% Iterate until convergence

for n = 1:inf % index

% Generate time span matrix (sec)

% Time interval

ti = te/40;

% Time span matrix

tspan = ts:ti:te;

[t,x] = ode15s('batch1',tspan,x_ss,opts); % Time and state outputs

% Parse out the state values

Ca(:,1) = x(:,1); % M

Cb(:,1) = x(:,2); % M

% Measure reactor performance

Yb = (Cb./(Ca + Cb))*100; % mol %

% Equilibrium yield calculation (mol %)

Yeq = exp(-delHovr/(R*TK))/(1+exp(-delHovr/(R*TK)))*100;

% Converge on equilibrium yield of B (mol %)

if abs(Yb(end) - Yeq) > 0.01 % Convergence criterion

te = te*2; % sec

ns = n; % Indicates that static catalysis criterion was not met

ns; % Display iteration count

clear t x Ca Cb Yb

else

% Stop static catalysis timer (sec)

toc

% Store static catalysis results

tsv(:,1) = t; % sec

Ca_star(:,1) = x(:,3); % gmol

Cb_star(:,1) = x(:,4); % gmol

% Remove static catalysis data and run

clear t x

break

end

end

% Begin dynamic catalysis

Page 42: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S18

% Initial conditions for dynamic catalysis

% Feed pressure (bar)

Pafb = rand*(Paf + Pbf);

Pbfb = rand*(Paf + Pbf);

if (Pafb + Pbfb) > (Paf + Pbf)

Pbfb = Paf + Pbf - Pafb;

end

% Feed concentration (M)

Cafb = Pafb/(Rg*TK);

Cbfb = Pbfb/(Rg*TK);

% Surface coverage (unitless)

Ca_star_b = rand*u;

Cb_star_b = rand*u;

if (Ca_star_b + Cb_star_b) > u

Cb_star_b = u - Ca_star_b;

end

x_ss(:,2) = [Cafb;Cbfb;Ca_star_b;Cb_star_b];

% Start dynamic catalysis timer (sec)

tic

% Iterate until convergence

for n = 1:inf % index

% Begin dynamic catalysis

for l = 2:Nosc % index

% Even numbered oscillations

if mod(l,2) == 0

% Operate at strong binding (eV)

delBEb = UR;

% Construct new time span matrix (sec)

% Final time

te = tau1;

% Time interval

ti = te/40;

% Time span matrix

tspan = ts:ti:te;

% Generate batch reactor data

[t,x] = ode15s('batch1',tspan,x_ss(:,l),opts); % Time and state outputs

% Parse out the state values

tsv(:,l) = t + tsv(end,l-1); % sec

Ca(:,l) = x(:,1); % M

Cb(:,l) = x(:,2); % M

Ca_star(:,l) = x(:,3); % gmol

Cb_star(:,l) = x(:,4); % gmol

% Initial conditions for the next oscillation

Page 43: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S19

x_ss(:,l+1) = [Ca(end,l);Cb(end,l);Ca_star(end,l);Cb_star(end,l)];

clear t x

% Odd numbered oscillations

else

% Operate at weak binding (eV)

delBEb = UL;

% Construct new time span matrix (sec)

% Final time

te = tau2;

% Time interval

ti = te/40;

% Time span matrix

tspan = ts:ti:te;

% Generate batch reactor data

[t,x] = ode15s('batch1',tspan,x_ss(:,l),opts); % Time and state outputs

% Parse out the state values

tsv(:,l) = t + tsv(end,l-1); % sec

Ca(:,l) = x(:,1); % M

Cb(:,l) = x(:,2); % M

Ca_star(:,l) = x(:,3); % gmol

Cb_star(:,l) = x(:,4); % gmol

% Initial conditions for the next oscillation

x_ss(:,l+1) = [Ca(end,l);Cb(end,l);Ca_star(end,l);Cb_star(end,l)];

clear t x

end

end

% End dynamic catalysis

% Calculate state values (gmol)

C_star = u - Ca_star - Cb_star;

% Convert units

Pa = Ca*Rg*TK; % bar

Pb = Cb*Rg*TK; % bar

Theta_A_star = Ca_star/u; % unitless

Theta_B_star = Cb_star/u; % unitless

Theta_star = C_star/u; % unitless

% Measure reactor performance

Yb = (Cb./(Ca + Cb))*100; % mol %

% Converge on steady state yield of B (mol %)

% Data at the end of the run

% Time data (sec)

tdiff = [tsv(:,end-1);tsv(:,end)];

% Yield data (mol %)

Ybdiff = [Yb(:,end-1);Yb(:,end)];

Page 44: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S20

% Data in the middle of the run

% Time data (sec)

tdiff2 = [tsv(:,(end-1)/2);tsv(:,(end-1)/2+1)];

% Yield data (mol %)

Ybdiff2 = [Yb(:,(end-1)/2);Yb(:,(end-1)/2+1)];

% Integrate yield of B over 1 oscillation period

% Preallocate integral matrix (sec*mol %)

Ybint = zeros(size(Ybdiff));

% Integration with midpoint Riemann sum

for m = 2:size(tdiff,1)

Ybint(m) = (tdiff(m) - tdiff(m-1))*mean([Ybdiff(m-1);Ybdiff(m)]);

End

% Preallocate integral matrix (sec*mol %)

Ybint2 = zeros(size(Ybdiff2));

% Integration with midpoint Riemann sum

for p = 2:size(tdiff2,1)

Ybint2(p) = (tdiff2(p) - tdiff2(p-1))*mean([Ybdiff2(p-1);Ybdiff2(p)]);

end

% Calculate time-averaged yield of B (mol %)

Ybavg = sum(Ybint)*fosc;

Ybavg2 = sum(Ybint2)*fosc;

% Check for convergence

if abs(Ybavg - Ybavg2) > 0.01 % Convergence criterion

Nosc = round(2*Nosc) + 1; % Adjust number of oscillations

nn = n; % Indicates that dynamic catalysis criterion was not met

nn; % Display iteration number

clear tsv(:,2:end) Ca(:,2:end) Cb(:,2:end) Ca_star(:,2:end) Cb_star(:,2:end) x_ss(:,3:end)

else

% Stop dynamic catalysis timer (sec)

toc

break

end

end

% Calculate ddG (kJ/gmol)

ddG = -(log(Ybavg/(100-Ybavg)) - log(Yeq/(100-Yeq)))*R*TK/1000;

% Calculate efficiency (%)

eff = abs(ddG/(delU*96.485))*100;

% Display final results

plot(tsv,Yb) % Visually check for convergence

Ybavg; % mol %

ddG; % kJ/gmol

eff; % %

Page 45: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S21

Dynamic Catalysis Batch Reactor Shell Code with Sinusoidal Waveform

% Remove prior data and runs

clear

clc

% Step test for models - CSTR and batch reactor

global Rg R TK Caf Cbf qdot w u delHovr alpha beta gamma delta BEb0 delBEb tau UR UL tdel

% Constants:

% Gas constant

Rg = 8.314459848e-2; % L-bar/K-gmol

R = 8.314459848; % J/K-gmol

% Reactor parameters:

% Temperature

Tc = 250; % deg C

TK = Tc + 273.15; % K

% Feed pressure (bar)

Paf = 100;

Pbf = 0;

% Feed concentration (M)

Caf = Paf/(Rg*TK);

Cbf = Pbf/(Rg*TK);

% Volumetric flowrate

qsdot = 50; % mL/min

qdot = qsdot/60000; % L/sec

% Steady state initial conditions for the states

Ca_ss = Caf; % M

Cb_ss = Cbf; % M

Ca_star_ss = 0; % gmol

Cb_star_ss = 0; % gmol

x_ss = [Ca_ss;Cb_ss;Ca_star_ss;Cb_star_ss];

% Steady state initial condition for the control

u_ss = 0; % gmol active sites

% Open loop step change

% Catalyst weight (mg)

w = 138;

% Active site density (gmol/g catalyst)

srho = 20e-6;

% Active site calculation (gmol)

u = (w/1000)*srho;

% Reaction parameters:

% Overall heat of reaction for A -> B (J/gmol)

delHovr = 0e3;

Page 46: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S22

% Brønsted-Evans-Polanyi relationship parameters

alpha = 0.6; % unitless

beta = 135e3; % J/gmol

% Binding energy relationship parameters

gamma = 2.0; % unitless

delta = 1.4; % eV

% Volcano plot parameter:

% x-axis zero point (eV)

BEb0 = delta + (delHovr/96.485e3);

% Volcano plot peak calculation

% Iterate binding energies of A and B (eV)

is = -delta;

ii = 0.01;

ie = delta;

% Preallocate BEb, TOF, and surface coverage matrices

je = ((ie - is)/ii + 1); % final index

delBEbv = zeros(round(je),1); % eV

TOFbvf = zeros(round(je),1); % 1/sec

Theta_A_starvf = zeros(round(je),1); % unitless

Theta_B_starvf = zeros(round(je),1); % unitless

% Begin iteration (eV)

for i = is:ii:ie

% Set binding energy of B (eV)

delBEb = i; % eV

% Iterate at each binding energy

for n = 1:inf % index

% Final Time (sec)

tfv = 5e100;

% Generate CSTR data

options = odeset('RelTol',1e-8,'AbsTol',1e-9);

[tv,xv] = ode15s('cstr1',[0 tfv],x_ss,options); % Time and state outputs

% Parse out the state values (M)

Cav = xv(:,1);

Cbv = xv(:,2);

% Measure reactor performance

Ybv = (Cbv./(Cav + Cbv))*100; % mol %

% Converge on C yield of B (mol %)

C = 1;

if abs(Ybv(end) - C) > 0.01 % Convergence criterion

qdot = qdot*Ybv(end)/C; % L/sec

nv = n; % Indicates that volcano criterion was not met

nv; % Display iteration count

clear tv xv Cav Cbv Ybv

else

% Compute matrix index

Page 47: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S23

j = ((i - is)/ii) + 1; % unitless

% Parse out state values (gmol)

Ca_starv = xv(:,3);

Cb_starv = xv(:,4);

% Measure reactor performance

TOFbv = Cbv*qdot/u; % 1/sec

% Convert units

Theta_A_starv = Ca_starv/u; % unitless

Theta_B_starv = Cb_starv/u; % unitless

% Store volcano plot results

delBEbv(round(j)) = i; % eV

TOFbvf(round(j)) = TOFbv(end); % 1/sec

Theta_A_starvf(round(j)) = Theta_A_starv(end); % unitless

Theta_B_starvf(round(j)) = Theta_B_starv(end); % unitless

clear tv xv Cav Cbv Ybv Ca_starv Cb_starv TOFbv Theta_A_starv Theta_B_starv

break

end

end

end

% Extract volcano plot peak information

peakTOF = max(TOFbvf); % 1/sec

% Preallocate relative binding energy matrix (eV)

peakdelBEb = zeros(round(je),1);

% Iterate relative binding energy of B (eV)

for k = 1:je % index

if abs(TOFbvf(k) - peakTOF) < peakTOF/100 % Convergence criterion

peakdelBEb(k) = delBEbv(k); % Store relative binding energy

end

end

% Calculate volcano plot peak position (eV)

peakdelBEb = mean(nonzeros(peakdelBEb));

% Display volcano plot peak results

peakdelBEb; % eV

peakTOF; % 1/sec

% Transition to dynamic catalysis

% Dynamic catalysis parameters

% Oscillation frequency

tau = 1e0; % sec

fosc = 1/tau; % Hz

% Number of oscillations (unitless)

Nosc = fosc;

Page 48: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S24

if Nosc < 11

Nosc = 11;

end

% Dynamic time span (sec)

tdyn = tau*Nosc*10;

% Oscillation endpoints (eV)

delBEb0 = peakdelBEb;

delU = 0.6;

UR = delBEb0 + delU/2.0;

UL = UR - delU;

% Equilibrium yield calculation (mol %)

Yeq = exp(-delHovr/(R*TK))/(1+exp(-delHovr/(R*TK)))*100;

% Begin dynamic catalysis

% Start dynamic catalysis timer (sec)

tic

% Batch reactor initial conditions

% Feed pressure (bar)

Pbfb = rand*(Paf + Pbf);

Pafb = Paf + Pbf - Pbfb;

% Feed concentration (M)

Cafb = Pafb/(Rg*TK);

Cbfb = Pbfb/(Rg*TK);

% Surface coverage (gmol)

Ca_star_ssb = rand*u;

Cb_star_ssb = rand*u;

if (Ca_star_ssb + Cb_star_ssb) > u

Cb_star_ssb = u - Ca_star_ssb;

End

% Initial condition array

x_ssb = [Cafb;Cbfb;Ca_star_ssb;Cb_star_ssb];

% Generate batch reactor data

tdel = rand*tau;

% Iterate until convergence

for n = 1:inf

[t,x] = ode15s('sin_batch1',[0 tdyn],x_ssb,options); % Time and state outputs

% Parse out the state values

Ca = x(:,1); % M

Cb = x(:,2); % M

% Measure reactor performance

Yb = (Cb./(Ca + Cb))*100; % mol %

Page 49: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S25

% Find convergence point

index = zeros(size(Yb));

for l = 1:size(Yb,1)

if abs(Yb(end)-Yb(l)) > 0.5

index(l) = l;

end

end

index = max(index);

if index > size(Yb,1)/2

tdyn = 2*tdyn;

else

break

end

end

% Stop dynamic catalysis timer (sec)

toc

% Parse out the state values

Ca_star = x(:,3); % gmol

Cb_star = x(:,4); % gmol

C_star = u - Ca_star - Cb_star; % gmol

% Convert units

Pa = Ca*Rg*TK; % bar

Pb = Cb*Rg*TK; % bar

Theta_A_star = Ca_star/u; % unitless

Theta_B_star = Cb_star/u; % unitless

Theta_star = C_star/u; % unitless

% Track binding energies

delBEbf = (((UR - UL)/2)*(cos(2*pi()*(t-tdel)/tau))) + ((UR + UL)/2);

% Display final results

plot(t,Yb) % Visually check for convergence

Page 50: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S26

Section S2. ODE Solver Selection and Justification

Matlab ODE solvers were screened for several different reactor types and catalytic systems under static

conditions or using a sinusoidal dynamic catalysis waveform. ODE45 is the recommended solver for

general use in Matlab, however, this solver failed to generate solutions for static and dynamic catalysis.

Static Catalysis

Stats for ode15s:

452 successful steps

58 failed attempts

1047 function evaluations

39 partial derivatives

131 LU decompositions

851 solutions of linear systems

Elapsed time is 0.159894 seconds.

Stats for ode23s:

2546 successful steps

1197 failed attempts

20218 function evaluations

2546 partial derivatives

3743 LU decompositions

11229 solutions of linear systems

Elapsed time is 1.949530 seconds.

Stats for ode23t:

1143 successful steps

8 failed attempts

2207 function evaluations

8 partial derivatives

106 LU decompositions

2166 solutions of linear systems

Elapsed time is 0.347212 seconds.

Stats for ode23tb:

1054 successful steps

8 failed attempts

2902 function evaluations

5 partial derivatives

99 LU decompositions

3934 solutions of linear systems

Elapsed time is 0.276370 seconds.

Stats for radau:

72 successful steps

72 failed attempts

1377 function evaluations

Page 51: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S27

147 LU decompositions

397 solutions of linear systems

Elapsed time is 0.210526 seconds.

ODE23t and ODE23tb have the least number of failed attempts, but Radau performed most efficiently in

terms of number of steps and ODE15s in terms of time to generate the solution. Therefore, ODE15s was

used throughout this manuscript to solve CSTR and Batch Reactor equations under static catalysis

conditions.

CSTR with Gamma < 1.0 and Sinusoidal Waveform

ODE23s, ODE23tb, and Radau did not converge on a stable sinusoidal CSTR solution after 8 h of

computational time.

Stats for ode15s:

18819 successful steps

3682 failed attempts

46463 function evaluations

1291 partial derivatives

5906 LU decompositions

40007 solutions of linear systems

Elapsed time is 7.945276 seconds.

Stats for ode23t:

51424 successful steps

2625 failed attempts

102764 function evaluations

1671 partial derivatives

8781 LU decompositions

94408 solutions of linear systems

Elapsed time is 17.107914 seconds.

ODE23t and has the least number of failed attempts, but ODE15s performed most efficiently in terms of

number of steps and time to generate the solution. Therefore, ODE15s was used throughout this

manuscript to solve CSTR and Batch Reactor equations under dynamic catalysis conditions.

Page 52: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S28

Section S3. Static Catalysis Time on Stream Data

CSTR at 1 % Yield of B

Figure S1. Example CSTR time on stream data with A) gas phase concentration of A and B and

B) amount of adsorbed species A* and B*. Conditions: 150 oC, 100 bar, and 1 % yield of B.

Batch Reactor at 50 % Yield of B

Figure S2. Example Batch Reactor time on stream data with A) gas phase concentration of A

and B and B) amount of adsorbed species A* and B*. Conditions: 150 oC, 100 bar, and

equilibrium yield of B.

Page 53: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S29

Section S4. Data from Heatmap Figures

Table S1. CSTR TOF (1/s) heatmap data for gamma of 0.5 with varying oscillation endpoints

and frequency. Conditions: 100 oC, 100 bar, and 1 % yield of B. Reaction parameters: ΔHovr of 0

kJ/mol, alpha of 0.6, beta of 102 kJ/mol, gamma of 0.5, and delta of 1.4 eV. Oscillation

amplitude of 1.03 eV.

0 -0.10 -0.21 -0.31 -0.41 -0.52 -0.62 -0.72 -0.82 -0.93 -1.03

0.01 0.0364 0.0256 0.0192 0.0176 0.0188 0.0252 0.0282 0.0281 0.0280 0.0280 0.0280

0.1 0.0364 0.0257 0.0200 0.0187 0.0209 0.0418 0.0644 0.0669 0.0669 0.0669 0.0669

1 0.0364 0 0.0199 0.0191 0.0315 0.1817 0.3434 0.3620 0.3629 0.3629 0.3629

10 0.0364 0 0 0 0.1385 1.629 3.294 3.486 3.495 3.495 3.495

100 0.0364 0 0 0 0.7139 9.540 16.98 17.65 17.68 17.68 17.68

1000 0.0364 0 0 0 1.020 13.88 24.02 24.84 24.88 24.88 24.88

10000 0.0364 0 0 0 0 14.51 25.00 25.83 25.87 25.87 25.87

100000 0.0364 0 0 0 0 14.55 25.06 25.89 25.93 25.93 25.93

Table S2. Batch reactor composition (mol % B) heatmap data for gamma of 2.0 with varying

oscillation endpoints and frequency. Conditions: 150 oC, 100 bar, and steady state yield of B.

Reaction parameters: ΔHovr of 0 kJ/mol, alpha of 0.6, beta of 102 kJ/mol, gamma of 2.0, and

delta of 1.4 eV. Oscillation amplitude of 0.5 eV.

-0.2 -0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

0.000001 50.00 50.00 50.00 50.00 50.01 50.01 50.02 50.01 50.00 50.00 50.00

0.00001 49.98 50.00 50.00 50.01 50.02 50.09 50.22 50.14 50.07 50.04 50.01

0.0001 49.99 50.00 50.00 50.01 50.06 50.66 52.61 52.73 51.24 50.44 50.21

0.001 49.99 50.00 50.00 50.03 50.49 55.91 69.00 69.14 63.25 57.27 53.31

0.01 49.98 49.95 50.01 50.31 54.79 87.24 93.69 92.80 88.42 81.79 71.86

0.1 49.99 49.92 50.10 53.06 91.80 99.02 99.18 98.79 98.29 96.97 94.09

1 49.17 49.26 50.44 77.73 99.65 99.84 99.79 99.61 99.52 99.34 98.61

10 26.00 35.62 59.77 96.98 99.78 99.95 99.94 99.91 99.88 99.88 99.90

100 18.43 27.60 66.32 97.21 99.79 99.96 99.97 99.97 99.97 99.97 99.97

Table S3. Batch reactor composition (mol % B) heatmap data for gamma of 2.0 with varying

oscillation endpoints and frequency. Conditions: 150 oC, 100 bar, and steady state yield of B.

Reaction parameters: ΔHovr of 0 kJ/mol, alpha of 0.6, beta of 102 kJ/mol, gamma of 2.0, and

delta of 1.4 eV. Oscillation amplitude of 1.0 eV.

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

0.000001 50.16 52.38 50.89 50.72 49.90 50.94 50.32 50.04 50.00 50.00 50.00

0.00001 49.34 49.79 50.03 49.40 51.90 59.85 54.01 50.76 50.10 50.01 50.00

0.0001 49.50 49.17 49.49 53.91 66.34 85.70 75.96 58.62 51.66 50.21 50.03

0.001 49.23 43.99 44.88 74.52 91.44 97.69 95.55 83.97 64.31 52.56 50.62

0.01 45.34 22.38 37.31 92.37 99.05 99.71 98.85 97.06 90.06 72.12 56.66

0.1 25.29 6.21 39.68 97.33 99.90 99.98 99.81 99.30 98.36 94.20 80.59

1 4.72 3.47 39.53 97.73 99.98 100.00 100.00 99.80 99.52 99.20 97.19

10 0.60 3.22 39.52 97.76 99.99 100.00 100.00 99.98 99.89 99.93 99.80

100 0.24 3.21 39.53 97.76 99.99 100.00 100.00 100.00 100.00 100.00 99.98

Page 54: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S30

Table S4. Batch reactor composition (mol % B) heatmap data for gamma of 2.0 with varying

oscillation endpoints and frequency. Conditions: 150 oC, 100 bar, and steady state yield of B.

Reaction parameters: ΔHovr of 0 kJ/mol, alpha of 0.6, beta of 102 kJ/mol, gamma of 2.0, and

delta of 1.4 eV. Oscillation amplitude of 1.5 eV.

0.0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4 2.7 3.0

0.000001 49.69 49.65 50.96 50.12 47.07 51.05 50.16 50.94 50.03 50.00 50.00

0.00001 49.62 49.56 59.79 53.98 50.37 47.82 52.05 59.91 50.75 50.04 50.00

0.0001 49.59 47.26 83.50 84.69 57.65 52.96 66.47 84.29 58.40 50.58 50.04

0.001 48.85 34.51 87.83 95.17 82.83 71.89 91.50 97.66 83.91 57.41 50.60

0.01 45.46 16.37 88.55 99.29 97.63 94.39 99.11 99.53 97.37 82.14 56.66

0.1 25.45 11.89 88.65 99.71 99.74 99.38 99.91 99.98 99.20 97.01 80.59

1 4.70 11.38 88.66 99.79 99.98 99.96 99.99 100.00 99.78 99.29 97.18

10 0.60 11.34 88.66 99.79 99.99 99.99 100.00 100.00 99.99 99.88 99.80

100 0.24 11.34 88.66 99.79 100.00 100.00 100.00 100.00 100.00 100.00 99.98

Table S5. Batch reactor composition (mol % B) heatmap data for gamma of 0.5 with varying

oscillation endpoints and frequency. Conditions: 150 oC, 100 bar, and steady state yield of B.

Reaction parameters: ΔHovr of 0 kJ/mol, alpha of 0.6, beta of 102 kJ/mol, gamma of 0.5, and

delta of 1.4 eV. Oscillation amplitude of 0.5 eV.

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50

0.000001 50.00 50.00 50.00 50.00 50.00 50.00 50.00 50.00 50.00 50.00 50.00

0.00001 50.00 50.00 50.00 50.00 49.99 50.00 50.00 49.99 49.99 50.00 50.00

0.0001 50.02 50.01 50.00 50.00 50.00 49.98 49.93 49.89 49.88 49.91 49.93

0.001 50.03 50.01 50.00 50.00 49.98 49.82 49.18 48.31 47.95 48.42 49.07

0.01 50.05 50.02 50.01 50.00 49.84 48.21 42.55 36.49 34.06 36.21 40.22

0.1 50.05 50.02 50.01 50.00 48.39 34.44 16.94 10.54 8.884 10.47 14.46

1 50.06 50.02 50.01 50.00 34.60 5.573 2.340 1.447 1.363 1.663 2.249

10 50.06 50.02 50.01 50.00 6.778 1.030 0.5733 0.5133 0.5231 0.5665 0.5999

100 73.78 75.94 74.63 49.99 5.182 0.7511 0.4453 0.4232 0.4231 0.4270 0.4275

Table S6. Batch reactor composition (mol % B) heatmap data for gamma of 0.5 with varying

oscillation endpoints and frequency. Conditions: 150 oC, 100 bar, and steady state yield of B.

Reaction parameters: ΔHovr of 0 kJ/mol, alpha of 0.6, beta of 102 kJ/mol, gamma of 0.5, and

delta of 1.4 eV. Oscillation amplitude of 1.0 eV.

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0.000001 50.23 50.09 50.01 50.00 50.00 50.00 50.00 49.99 50.00 50.00 50.00

0.00001 50.38 50.34 50.18 50.04 50.00 49.98 49.93 49.83 49.95 49.99 50.00

0.0001 50.39 50.39 50.42 50.53 49.92 49.67 48.74 47.04 49.16 49.80 49.94

0.001 50.40 50.45 51.57 55.31 48.64 45.23 37.92 29.56 40.77 46.69 49.03

0.01 50.45 51.25 61.15 74.50 40.67 23.71 11.94 6.554 15.18 28.23 39.89

0.1 50.98 58.15 86.67 88.18 29.33 4.267 1.654 1.140 2.470 6.000 14.06

1 55.70 82.71 97.77 90.43 26.52 0.7109 0.1376 0.1313 0.6546 1.166 2.177

10 77.49 97.48 99.25 90.64 26.29 0.4056 0.0190 0.0060 0.1439 0.3561 0.3556

Page 55: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S31

Table S7. Batch reactor composition (mol % B) heatmap data for gamma of 0.5 with varying

oscillation endpoints and frequency. Conditions: 150 oC, 100 bar, and steady state yield of B.

Reaction parameters: ΔHovr of 0 kJ/mol, alpha of 0.6, beta of 102 kJ/mol, gamma of 0.5, and

delta of 1.4 eV. Oscillation amplitude of 1.5 eV.

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0.000001 50.23 50.09 50.01 50.00 50.00 50.00 50.00 49.99 50.00 50.00 50.00

0.00001 50.38 50.34 50.18 50.04 50.00 49.98 49.93 49.83 49.95 49.99 50.00

0.0001 50.39 50.39 50.42 50.53 49.92 49.67 48.74 47.04 49.16 49.80 49.94

0.001 50.40 50.45 51.57 55.31 48.64 45.23 37.92 29.56 40.77 46.69 49.03

0.01 50.45 51.25 61.15 74.50 40.67 23.71 11.94 6.554 15.18 28.23 39.89

0.1 50.98 58.15 86.67 88.18 29.33 4.267 1.654 1.140 2.470 6.000 14.06

1 55.70 82.71 97.77 90.43 26.52 0.7109 0.1376 0.1313 0.6546 1.166 2.177

10 77.49 97.48 99.25 90.64 26.29 0.4056 0.0190 0.0060 0.1439 0.3561 0.3556

Table S8. Batch reactor efficiency (%) heatmap data for gamma of 0.5 with varying oscillation

endpoints and amplitudes. Conditions: 150 oC, 100 bar, and steady state yield of B. Reaction

parameters: ΔHovr of 0 kJ/mol, alpha of 0.6, beta of 102 kJ/mol, gamma of 0.5, and delta of 1.4

eV. Oscillation frequency of 1 Hz.

0.00 0.15 0.30 0.45 0.60 0.75 0.90 1.05 1.20 1.35 1.50

0.00 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000

0.15 -0.0002 -0.3224 -7.612 -12.61 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000

0.30 0.0063 -0.2547 -10.52 -13.65 -7.409 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000

0.45 0.0957 -0.1519 -13.02 -14.85 -9.846 -4.949 0.0000 0.0000 0.0000 0.0000 0.0000

0.60 0.7986 1.962 -12.27 -13.79 -10.44 -7.394 -3.712 0.0000 0.0000 0.0000 0.0000

0.75 1.102 10.88 -3.757 -15.71 -11.01 -9.258 -5.915 -2.970 0.0000 0.0000 0.0000

0.90 0.9411 11.04 6.624 -10.38 -13.63 -9.247 -7.573 -4.930 -2.482 0.0000 0.0000

1.05 0.7954 9.554 7.996 -3.502 -10.48 -11.81 -8.199 -6.612 -4.215 -2.121 0.0000

1.20 0.6960 8.368 7.056 -2.714 -7.385 -9.223 -10.07 -7.027 -5.786 -3.695 -1.856

1.35 0.6186 7.438 6.273 -2.407 -7.558 -6.842 -8.202 -8.949 -5.629 -5.143 -3.283

1.50 0.5568 6.694 5.646 -2.167 -7.130 -8.343 -6.162 -7.360 -8.183 -5.066 -4.629

Page 56: Catalytic Resonance Theory: SuperVolcanoes, Catalytic

Ardagh, et al. Supplementary Information Page S32

Section S5. Binding Energy Derivation

In this section, we demonstrate that the binding energies of A and B can be calculated at any

point by using the definitions of ΔHovr, gamma, and delta.

Definitions for gamma and delta are summarized in Equations S1 and S2:

𝛾 ≡∆𝐵𝐸𝐵

∆𝐵𝐸𝐴 (S1)

𝑤ℎ𝑒𝑛 𝐵𝐸𝐴 = 𝛿, 𝐵𝐸𝐵 = 𝛿 + ∆𝐻𝑜𝑣𝑟 (S2)

Equation S1 can be integrated with an indefinite integral to obtain Equation S3:

𝐵𝐸𝐵 = 𝛾𝐵𝐸𝐴 + 𝐶 (S3)

After plugging Equation S2 into Equation S3, the constant of integration is obtained:

𝐶 = (1 − 𝛾)𝛿 + ∆𝐻𝑜𝑣𝑟 (S4)

Equation S4 can therefore be substituted into Equation S3 to define a relationship between the

binding energies of A and B at any point.

𝐵𝐸𝐵 = 𝛾𝐵𝐸𝐴 + (1 − 𝛾)𝛿 + ∆𝐻𝑜𝑣𝑟 (S5)

Alternatively, Equation S5 may be rearranged to calculate the binding energy of A from a known

binding energy of B.

𝐵𝐸𝐴 = (𝐵𝐸𝐵 − (1 − 𝛾)𝛿 − ∆𝐻𝑜𝑣𝑟)/𝛾 (S6)