Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

Supporting Information

Two-Dimensional Chemiresistive Covalent Organic Framework with High Intrinsic


Conductivity
Zheng Meng,1 Robert M. Stolz,1 Katherine A. Mirica1*
1
Department of Chemistry, Burke Laboratory, Dartmouth College, Hanover, New
Hampshire 03755

*E-mail: Katherine.A.Mirica@dartmouth.edu

Contents
1. Materials and Methods ......................................................................................................... 2
2. Synthesis of COF-DC-8 ......................................................................................................... 3
2.1. Synthesis of (2,3,9,10,16,17,23,24-octaaminophthalocyaninato) nickel(II) NiOAPc ..... 3
2.2. Synthesis of COF-DC-8 .................................................................................................. 8
3. FTIR Spectroscopy .............................................................................................................. 11
4. X-ray Photoelectron Spectroscopy ..................................................................................... 12
5. Elemental Analysis .............................................................................................................. 14
6. Structural Analysis by Computational Study and Powder X-ray diffraction ............... 15
7. Computational Study of Electronic Properties ................................................................. 18
8. SEM and TEM ..................................................................................................................... 19
9. Brunauer–Emmett–Teller (BET) Analysis ....................................................................... 21
10. Thermal Gravimetric Analyses (TGA) .......................................................................... 21
11. Conductivity Measurement and Activation Energy Determination ........................... 22
12. Gas Sensing Experiments................................................................................................ 24
12.1. Fabrication of Gas Sensors ........................................................................................ 24
12.2. Sensing Experiments ................................................................................................. 25
12.2.1. NH3 Sensing .......................................................................................................... 27
12.2.2. H2S Sensing ........................................................................................................... 28
12.2.3. NO Sensing............................................................................................................ 30
12.2.4. NO2 Sensing .......................................................................................................... 32
12.2.5. Analysis of the Kinetic of Sensing Response ........................................................ 33
12.2.6. Recovery of the Sensing Performance by Thermal and Solvent Activation ......... 37
12.3. Sensing Mechanism by XPS and EPR ...................................................................... 38
13. References ........................................................................................................................ 42

S1
1. Materials and Methods
Nickel(II) chloride (CAS N.O.: 7718-54-9), cobalt(II) chloride (CAS N.O.: 7646-79-9),
copper cyanide (CAS N.O.: 544-92-3), 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU, CAS N.O.:
6674-22-2), were purchased form Sigma Aldrich. Methanol (CAS N.O.: 67-56-1), dichloromethane
(CAS N.O.: 75-09-2), acetone (CAS N.O.: 67-64-1), N, N-dimethylacetamide (CAS N.O.: 127-19-
5), acetic acid (CAS N.O.: 64-19-7), N-methyl-2-pyrrolidone (NMP, CAS N.O.: 872-50-4),
petroleum ether (CAS N.O.: 8032-32-4), dimethylacetamide (CAS N.O.: 127.19-5), 1,2-
dichlorobenzene (DCB, CAS N.O.: 95-50-1) and ethyl acetate (CAS N.O.: 141-78-6) were
purchased from BDH Chemicals. Pyrene-4,5,9,10-tetraone (TOPyr) was synthesized by following
the literature procedure.S1 The chemical shifts (δ) were expressed in ppm with internal standard
tetramethylsilane (TMS) and solvent signals as internal references, and J values were given in Hz.
Standard abbreviations indicating multiplicity were used as follows: s (singlet), br (broad), d
(doublet), t (triplet), q (quartet), m (multiplet).
Scanning Electron Microscopy (SEM) was performed using a Hitachi TM3000 SEM
(Tokyo, Japan) equipped for X-ray microanalysis with a Bruker Edax light element Si(Li) detector
(Billerica, MA). Transmission electron microscopy was carried out at a Tecnai F20ST FEG TEM
instrument. Powder X-ray diffraction (PXRD) measurements were performed with a Rigaku sixth
generation MiniFlex X-ray diffractometer with a 600 W (40 kV, 15 mA) CuK (= 1.54 Å)
radiation source. Nitrogen adsorption experiments were performed with ASAP Plus 2020
(Mircromeritics, Norcross, Georgia) instrument. Elemental analyses were performed by Atlantic
Microlab, inc. X-ray photoelectron spectroscopy (XPS) experiments were conducted on Kratos
Analytical AXIS Supra X-ray Photoelectron Spectrometer under ultrahigh vacuum (base pressure
10-7 Torr). The measurement chamber was equipped with a monochromatic Al (Kα) X-ray source.
Both survey and high-resolution spectra were obtained using a beam diameter of 200 μm. Electron
paramagnetic resonance (EPR) spectra were collected on a Bruker BioSpin Gmbh spectrometer
equipped with a standard mode cavity.

S2
2. Synthesis of COF-DC-8
2.1. Synthesis of (2,3,9,10,16,17,23,24-octaaminophthalocyaninato) nickel(II)
NiOAPc

Synthesis of N,N'-(4,5-dicyano-1,2-phenylene)bis(4-methylbenzenesulfonamide) 4.
The starting material N,N'-(4,5-dibromo-1,2-phenylene)bis(4-methylbenzenesulfonamide) 3, as
shown in the above synthetic scheme, was synthesized from o-phylenediamine 1 in two steps using
literature procedures.S2 The synthetic and work-up procedures for N,N'-(4,5-dicyano-1,2-
phenylene)bis(4-methylbenzenesulfonamide) 4 were modified from the previously reported
procedures.S3 A mixture of 50 g (87.1 mmol) of N,N'-(4,5-dibromo-1,2-phenylene)bis(4-
methylbenzenesulfonamide), 23.5 g (262.5 mmol, 3 e.q.) of CuCN and 240 mL of anhydrous DMF
were heated to 120 °C for 48 h under N2 atmosphere.
After the reaction was cooled to room temperature, the resulting dark brown mixture was
mixed with aqueous NH3•H2O (25%~28%, 600 mL) and air was bubbled through the solution for
24 h. A brown precipitate was collected by filtration and washed with water until the filtrate became
colorless. The light green-yellow solid was dried and then dispersed into a mixture of
dichloromethane (1 L) and acetic acid (100 mL). The mixture was extracted by EDTA•2Na aqueous
solution (8 g in 300 mL H2O) for 2 times to remove Cu2+ residues. The aqueous phase was separated
and further extracted with 500 mL of dichloromethane. The organic phase was combined and
filtered through a fine fritted funnel to obtain a transparent and slightly brown solution. This organic
solution was collected and evaporated to give a light yellow solid. The solid obtained was

S3
recrystallized from 300 mL of acetic acid to give yellows crystals, which were collected by
filtration. These yellow crystals were further recrystallized from 200 mL of ethanol to improve the
purity and give the desired product 4 as light yellow crystals (28 g, yield 69%). 1H NMR (600 MHz,
298K, DMSO-d6) δ = 7.67 (d, J = 8.3 Hz, 2H), 7.58 (s, 1H), 7.36 (d, J = 8.1 Hz, 2H), 2.37 (s, 4H).
13
C NMR (600 MHz, 298K, DMSO-d6) δ=144.5, 136.5, 134.4, 130.4, 127.3, 124.3, 116.1, 109.6,
21.5. The characterization data match the literature.S3b

Figure S1. 1H NMR (600 M, 298 K, CDCl3) spectrum of 4.

S4
Figure S2. 13C NMR (150 M, 298 K, CDCl3) spectrum of 4.

Synthesis of 2,3,9,10,16,17,23,24-octatosylaminophthalocyanine Ni(II) 5. The synthesis


of the 2,3,9,10,16,17,23,24-octatosylaminophthalocyanine Ni(II) was adapted from the literature.S4
To a suspension of 4.01 g (8.6 mmol) of N,N'-(4,5-dicyano-1,2-phenylene)bis(4-
methylbenzenesulfonamide) and 1.12 g of anhydrous NiCl2 (8.6 mol, 1.0 eq) in n-hexanol (12 mL)
in a glass pressure vessel, 2 mL of DBU (1,8-Diazabicyclo[5.4.0]undec-7-ene) was added. The
system was purged with N2 for 10 minutes and then sealed. The reaction vessel was then heated at
160 °C under N2 for 36 hours. Upon cooling to room temperature, the reaction mixture was
dissolved in 250 mL mixed solvent of acetic acid and CH2Cl2 (v/v=1:5). The organic extract was
washed with water (200 mL  3) and then evaporated to dryness to generate an emerald-colored
solid. The residual high boiling point hexanol remaining in the solid was removed by repeatedly
dispersing the solid in 40 mL of CH2Cl2/petroleum ether mixture (v/v=1:3) and decanting the
brownish solvent. The emerald-colored crude product was further purified by column
chromatography (silica gel, eluent: CH2Cl2/MeOH=50/1) to give 5 as a green solid (1.27 g, yield
67%). 1H NMR (600 MHz, 298K, CDCl3) δ = 8.79 (s, 4H), 8.64 (s, 4H), 8.03 (s, 4H), 7.98 (d, J =
8.1 Hz, 8H), 7.87 (s, 4H), 7.74 (d, J = 8.0 Hz, 8H), 7.07 (d, J = 7.9 Hz, 8H), 6.66 (d, J = 7.9 Hz,
8H), 2.87 (s, 12H), 2.10 (s, 12H). The characterization data matched the literature report.S3b

S5
Figure S3. 1H NMR (600 M, 298 K, CDCl3) spectrum of 5.

Synthesis of NiOAPc. To a 50 mL round bottom flask charged with 1.0 g (0.51 mmol) of
2,3,9,10,16,17,23,24-octatosylaminophthalocyanine Ni(II), 3 mL of deionized water and 30 mL of
concentrated H2SO4 were successively added. The mixture was then heated at 110 °C for about 1.5
hour. The reaction was then cooled to room temperature, and the reaction mixture was poured into
ice-water (50 mL). The green precipitate was collected by centrifugation and the solid was washed
thoroughly with deionized water (20 mL  3), 10% NaOH (20 mL  3), deionized water (20 mL 
3), and acetone (20 mL  3) sequentially with the help of a vortex. The product NiOAPc was
obtained as a purple to black powder (327 mg, 93%). 1H NMR (600 MHz, DMSO-d6): δ = 8.36 (s,
8H), 5.83 (br, 16H); 13C NMR (150 MHz, 298K, DMSO-d6): δ = 145.1, 138.9, 129.8, 129.3, 105.5.

S6
Figure S4. 1H NMR (600 M, 298 K, DMSO-d6) spectrum of NiOAPc.

Figure S5. 13C NMR (150 M, 298 K, DMSO-d6) spectrum of NiOAPc.

S7
2.2. Synthesis of COF-DC-8

Different synthetic conditions, including variation of the solvent system, concentration of


ligand, and the use of acid, temperature and reaction time were explored for the synthesis of COF-
DC-8 (Table S1). The crystallinity of product was determined by PXRD. It was found that using a
mixed solvent (v/v=1/1) of 1,2-dichlorobenzene (DCB) and dimethylacetamide (DMAC) under an
elevated temperature in the presence of sulfuric acid gave the product with good crystallinity (entry
9 in Table S1). Efforts to optimize reaction conditions also yielded several alternative approaches
for accessing COF-DC-8, including the use of an aqueous acetic acid solution as the catalyst in a
mixed solvent of DMAC and DCB (entry 14 in Table S1), and the use of NMP as the solvent with
sulfuric acid or aqueous solution of acetic acid (entry 17,18 in Table S1). The experimental results
showed that the use of acid (entry 8 vs entry 9, entry 16 vs entry 17) and the use of a relatively low
concentration of the building blocks (entry 7 vs entry 8, entry 14 vs entry 15, entry 1 vs entry 17)
were beneficial for the improvement of crystallinity.

S8
Table S1. The optimization of synthetic conditions for COF-DC-8.

[NiOAPc] Temperature
Entry solvent additives time
/mM /°C

1 NMP 15 H2SO4 (9 eq) 175 3d

2 NMP 15 -- 202 3d

3 DMSO 3 -- 180 5d

4 DCB/DMAC (1/1) 3 TfOH (9 eq) 202 10 d

DCB/DMAC/6 M HOAc
5 3 -- 150 3d
(1/1/0.25)

6 pyridine 5 -- 120 5d

7 DCB/DMAC (1/1) 15 -- 202 5d

8 DCB/DMAC (1/1) 2 -- 202 10 d

9 DCB/DMAC (1/1) 2 H2SO4 (9 eq) 202 10 d

1,2,4-triclorobenzene/DMAC 10 h
10 15 H2SO4 (9 eq) 240
(1/1) (MW)

11 pyridine/DMAC (1/1) 2 -- 160 7d

DCB/DMAC/H2O
12 2 H2SO4 (9 eq) 200 7d
(1/1/0.25)
DCB/DMAC/H2O H2SO4 (9 eq),
13 2 200 7d
(1/1/0.25) urea (9 eq)
DCB/DMAC/6 M HOAc
14 2 -- 200 7d
(1/1/0.25)
DCB/DMAC/6 M HOAc 10 h
15 15 -- 160
(1/1/0.25) (MW)

16 NMP 2 -- 200 7d

17 NMP 2 H2SO4 (9 eq) 200 7d

18 NMP/6 M HOAc (1/0.25) 2 -- 200 7d

S9
Figure S6. PXRD of the materials obtained under synthetic conditions demonstrated in Table S1.

S10
Synthesis of COF-DC-8 under optimized conditions. To a 25 mL Teflon chamber
charged with 20.7 mg (0.03 mmol) of NiOAPc and 15.7 mg (0.6 mmol) of TOPyr, 10 mL of
dimethylacetamide, 10 mL of o-dichlorobenzene, and 1 drop (~27.1 mg, 9.0 e.q.) of concentrated
H2SO4 were successively added. The mixture was then sonicated for 1-2 minutes until NiOAPc
and TOPyr were completely dissolved. The solution was then degassed with N2 for 10 minutes in
a teflon chamber, which was then capped under N2 and was heated in an autoclave vessel in the
oven at 202 °C for 10 days. After the reaction was cooled to room temperature, the dark green solid
was collected by centrifugation and was then washed with acetone (20 mL × 2), N,N-
dimethylformamide (20 mL × 2), and acetone (20 mL × 3) with the help of a vortex. The solid was
then transferred to a vacuum (20 mTorr) oven at 65 °C and dried for 24 hours to give COF-DC-8
as dark green powder (22 mg, yield 69%).

Synthesis of COF-DC-8 under microwave (MW) heating. To a 10 mL glass vessel


(Discover SP Vessel) with 20.7 mg (0.03 mmol) of NiOAPc and 15.7 mg (0.6 mmol) of TOPyr, 1
mL of dimethylacetamide, 1 mL of 1,2,4-triclorobenzene, and 1 drop (~27.1 mg, 9.0 e.q.) of
concentrated H2SO4 were successively added. The mixture was then sonicated for 1-2 minutes,
degassed by N2 for 5 minutes and the vessel was then capped with a silicon septum under N2. The
glassware was transferred to a microwave synthesizer (Discover® SP). The temperature was set at
240 °C for 10 hours. The work-up process is the same with the above procedure.

3. FTIR Spectroscopy
Infrared spectra were collected using a JASCO model FT IR-6100 Fourier transform
infrared spectrophotometer. The material was mixed with KBr power in a mass ratio of ~1:50 and
then pressed into a pellet in a die with a diameter of 8 mm under nonspecific pressures. TOPyr and
NiOAPc were directly used after they were isolated as described in section 2.1. COF-DC-8 was
used after it was further dried in vacuum desiccator equipped with an oil pump for 24 hours under
room temperature after synthesis.

S11
Figure S7. Comparison of FTIR spectra of TOPyr, NiOAPc, and COF-DC-8.

4. X-ray Photoelectron Spectroscopy


X-ray photoelectron spectroscopy (XPS) experiments were conducted on a Kratos Analytical
AXIS Supra X-ray Photoelectron Spectrometer under ultrahigh vacuum (base pressure 10-7 Torr).
The measurement chamber was equipped with a monochromatic Al (Kα) X-ray source. The as-
synthesized COF-DC-8 was dried in a vacuum desiccator equipped with an oil pump for 24 hours
under room temperature. About 6 mg of COF powder was put into a vial and degassed with N2 (1
L/min) for 1 h, which was were stored under N2 until needed. The material was mounted by lightly
pressing onto copper tape which was mounted on a Dual-Height (Kratos) sample holder. A survey
spectrum was obtained from 0 eV–1400 eV to obtain elemental surface composition. High
resolution spectra were then obtained at energy regions specific to elements observed in the survey
spectrum (C 1s, N 1s, O 1s, Ni 2p 3/2). XPS was performed under a pressure of ~10 −7 Torr. Both
survey and high-resolution spectra were obtained using a beam diameter of 200 μm.

S12
Figure S8. XPS survey spectrum of COF-DC-8 showing the presence of C, N, O, and Ni elements.

Figure S9. Deconvoluted XPS spectra of COF-DC-8: (a) C 1s, (b) N 1s, (c) O 1s, and (c) Ni 2p.
The deconvoluted peaks are color coded and their corresponding ratios are given.

S13
XPS spectra revealed the exclusive presence of C, N, O, and Ni in COF-DC-8 in weight
percentages of 76.77%, 15.81%, 4.36%, and 3.06%, respectively. Deconvolution of high-resolution
scans of the C 1s spectrum showed emission lines with binding energies of 284.6, 285.9, 287.8 and
289.9 eV, which were ascribed to the C=C, C=N, C=O, and π-π* species, respectively. The
deconvolution of N1s spectrum showed mainly two peaks at binding energies of 398.6 and 400.2
eV, which were ascribed to C=N without and with coordination to Ni, respectively. The higher ratio
(64:24) of the uncoordinated C=N compared to the coordinated C=N indicated the formation of
new C=N bonds during the COF synthesis, because in the building block NiOAPc, the theoretical
uncoordinated/coordinated C=N is 1:1.S5 The small amount of O content may be due to the
existence of the unreacted C=O groups at sheet edges, or entrapment of other oxygen containing
species within the pores of the COF, such as acetone, DMAC, O2, or water. Studies have shown
that the unreacted C=O groups can be removed using an annealing procedure performed at high
temperature.S6 The high-resolution scans of the Ni 2p showed a typical spectrum of the
phthalocyanine bound Ni2+, in which the Ni 2p3/2 and Ni 2p1/1 satellite peaks are usually very
weak or missing.S7

5. Elemental Analysis
After COF-DC-8 was synthesized as described in section 2.1, it was further dried in vacuum
desiccator equipped with an oil pump for 24 hours under room temperature in preparation for
elemental analysis. Elemental analyses, including C, H, N, were performed by Atlantic Microlab
inc. using a combustion method by automatic analyzers. The metal content was analyzed by ICP-
MS. The results are listed below.
Table S2. Elemental analysis of COF-DC-8.
Sample Element Theoretical Found
C 71.73 65.60
H 1.88 3.41
COF-DC-8
N 20.91 17.32
Ni 5.48 4.58
Although absolute content of C, N, Ni were slightly lower than the theoretical values, the
C:N:Ni ratio matched the theoretical ratio very well. A significant high H content was observed. As
the small amount (4.36%, see section 4) of O was also observed by XPS, the discrepancy between
the theoretical and observed values in elemental analysis could be from the existence of the
unreacted C=O and –NH2 groups at sheet edges. The demonstration of the existence of the
unreacted C=O and –NH2 was also observed in the C2N COF synthesized by a similar condensation

S14
reaction, and an annealing procedure at high temperature could remove those groups.6 The TGA
profile of COF-DC-8 revealed a ~4% weight loss before 189 °C (see Figure S20 in section 10),
potentially due to the loss of small-molecule volatiles (acetone, H2O, or DMAC) included in the
pores, which could also contribute to the discrepancy.

6. Structural Analysis by Computational Study and Powder


X-ray diffraction
To build the crystal structure of COF-DC-8, a model molecule, which was derived from two
phthalocyanine fragments and one pyrene-4,5,9,10-tetraone fragment, was built. The chemical
structure is shown in Figure S10a. The molecular model was optimized using B3LYP density-
functional theory method with a basis set of 6-21G in Gaussian 09. The optimized structure (Figure
S10b) of the model molecule was imported into Materials Studio. The initial crystal structure was
constructed with a unit cell with the space group P1 with lattice parameters of α=β=γ=90 °. The a
and b values were estimated from the meal-metal distance in the model molecule. The c value (3.34
A) was estimated according to the distance between the metal center of phthalocyanine unit and
was chosen from the values calculated from the PXRD. The symmetry of the initial structure was
then found and imposed to a higher symmetry of P4/MMM with a threshold of 0.01 Å (fine level).

Figure S10. The (a) chemical structure and (b) optimized ball-stick structure of the model
molecule used for the construction of the initial COF-DC-8 structure.
The geometry optimization with cell parameters was performed using the CASTEP module
in Materials Studio with Perdew−Burke−Ernzerhof (PBE) generalized gradient approximation
(GGA). The cut-off energy was set at 300.0 eV. The convergence tolerance was set at 10-5 eV/atom
and the max force was set to 0.03 eV A-1. The Brillouin zones were sampled using a 1 × 1 × 8 k-
point mesh in the Monkhorst–Pack scheme. Calculation of the simulated powder diffraction pattern
was performed by Materials Studio Reflex Plus Module. The optimized crystal structure of the
COF-DC-8 with lattice parameters is shown in Figure S11. Modeling of the staggered structure
was performed in a similar manner, but with the space group I4/MMM (Figure S12).

S15
Figure S11. Views from c and b axes of modeled COF-DC-8 crystal structure with eclipsed
packing mode.

Figure S12. Views from c and b axes of modeled COF-DC-8 crystal structure with staggered
packing mode.

PXRD data were collected using a Rigaku sixth generation MiniFlex X-ray diffractometer.
Cu Kα radiation (λ = 1.5406 Å; 600 W, 40 kV, 15 mA) was focused using a planar Göbel mirror
riding the Kα line.

S16
Figure S13. PXRD spectra for TOPyr, NiOAPc, and experimental (blue line, entry 9 in Figure
S6) and simulated models in eclipsed (black line) and staggered packing (grey line) of COF-DC-
8.

Figure S14. Representation of the planes [100], [110] and [001] in the crystal structure of COF-
DC-8.

S17
To check the chemical stability of the COF, about 15 mg of COF-DC-8 sample was
immersed into 5 mL of 12 M HCl and another 15 mg of COF-DC-8 was immersed in 14 M NaOH
each for 24 hours under room temperature. The immersed COF samples were then filtrated, washed
with water (10 mL  5) and acetone (10 mL  3), and transferred into a vacuum oven set at 65 °C
and kept for 1 hour. The samples were then characterized by PXRD.

Figure S15. PXRD pattern of (a) as synthesized COF-DC-8, (b) COF-DC-8 after being immersed
in 12 M HCl for 24 hours, and (c) COF-DC-8 after being immersed in 14 M NaOH for 24 hours.

7. Computational Study of Electronic Properties


For computation of electronic properties of the COF-DC-8, including band structure, density
of states, the functional GGA with PBE was employed with an energy cutoff set at 530 eV. The
SCF tolerance is set at ultra-fine level at 5.0×10-7 eV/atom. The Brillouin zones were sampled using
a 2 × 2 × 4 k-point mesh in the Monkhorst-Pack scheme.

S18
Figure S16. (a) Calculated electronic band structure (left side) and density of state (right side) for
COF-DC-8 and (b) the corresponding first Brillouin zone and high-symmetry K-points.
Electronic band structure showed that Dirac bands crossed the Fermi level in both A−M, Γ−Z,
and R-X directions and exhibited wide band dispersions of more than 0.5 eV, suggesting the
intrinsically conductive nature of COF-DC-8 (Figure S16a). The density of states analysis showed
that, compared with H and Ni atoms, the p orbitals from C and N atoms contribute more
significantly to the Dirac bands. The high symmetry points in the first Brillouin zone showed that
the Dirac bands crossed the Fermi level through the out-of-plane directions (Figure S16b); while,
in the in-plane directions, including Γ−M, Z−R, Γ−M and Z-A, moderate band gaps ranging from
0.5−1.0 eV were observed. These above analysis suggested that the charge transport in COF-DC-
8 is anisotropic and that a significant mechanism contributing to the bulk conductivity may be
through pathways along the c-axis.S8 This through space charge-transport property is consistent
with those found in other phthalocyanine-S9 and porphyrin-based COFs,S10 that exhibited high
carrier mobility along the direction of the stacking due to the formation of periodic π-columns.

8. SEM and TEM


Scanning electron microscopy of COF-DC-8 was obtained using a Hitachi TM3000 SEM.
The powdered material was pressed onto carbon conductive tape that was attached to an aluminum
sample plate. The images were taken at a 10 mm working distance using a 15 kV beamline in a 10-
6
torr vacuum chamber.
Transmission electron microscopy was carried out with a Tecnai F20ST FEG TEM
instrument. The sample was prepared by drop casting an acetone suspension (0.5 mg in 5 mL) of
the COF onto a copper grid (300 mesh, 3.0 mm O.D). An operating voltage of 120 kV was used for

S19
imaging.

Figure S17. SEM images at different magnifications of COF-DC-8 synthesized using condition
listed in entry 9 in Table S1.

Figure S18. TEM images of COF-DC-8 synthesized using condition listed in entry 9 in Table
S1.

SEM revealed the presence of rod-shaped crystallites, whose lengths can range from hundreds
of nanometers to several micrometers and widths expand up to hundreds of nanometers (Figure
S17, S18). SEM also showed the existence a small number of objects with unspecified morphology
(Figure S17). TEM image provided visualization of a layered morphology (Figure S18). Regular
lines with a spacing of ~2.2 nm in TEM were observed. However, a closer observation of the regular
nanostructure of COF-DC-8 was challenging because the COF specimen showed instability to
electron beam under high magnification, as also found for other types of framework materials.S11

S20
9. Brunauer–Emmett–Teller (BET) Analysis
In order to assess the porosity of the COF, gas adsorption measurements were performed on
an ASAP Plus 2020 (Mircromeritics, Norcross, Georgia) with N2 at 77K. To remove the high
boiling point solvent used in the synthesis from the COFs, the samples were soaked in acetone for
2 days, during which the acetone was changed every 12 hours. The samples were then dried in the
oven under vacuum (20 mTorr, 65°C for 24 hours). Before gas adsorption measurements, the
samples were degassed under vacuum at 120 °C for 18 hours. For BET calculations, a full isotherm
with a fitting range of 0 to 0.3 P/P0 was used. The Brunauer-Emmett-Teller (BET) surface area for
the COF was 360 m2/g.

Figure S19. (a) Nitrogen sorption curves (filled circles: adsorption, open circles: desorption,
STP=standard temperature pressure), and the inset in (a) is pore size distribution. (b) t-Plot for
nitrogen adsorbed at 77 K by activated COF-DC-8. The BET surface area by the gas adsorption
analysis is 360 m2/g. The external surface area is 124 m2/g and the micropore area is 236 m2/g.

10. Thermal Gravimetric Analyses (TGA)


Thermal gravimetric analysis was performed using a TA Instruments TGA Q150 with a 20
o
C/min ramp from 40 oC to 900 oC under N2. TGA profile of COF-DC-8 revealed a ~4% weight
loss before 189 °C (Figure S20), potentially due to the loss of small-molecule volatiles (acetone,
H2O, or DMAC) included in the pores. With the temperature raised to 510 oC, only a 10 % mass
loss was observed, which suggests good thermal stability of the COF. We noticed that this 10%
mass loss at 510 oC was much lower than structurally related nickel phthalocyanine-S12 and copper
phthalocyanine-S13 based MOFs, which showed 30% to 50% weight loss before 500 oC.

S21
Figure S20. TGA curves of TOPyr, NiOAPc, and COF-DC-8.

11. Conductivity Measurement and Activation Energy


Determination
To make a pressed pellet, ~40 mg of COF sample was put into a 6 mm inner-diameter split
sleeve pressing die and pressed for 5 min under a pressure of 1000 psi. A Singatone tungsten carbide
four-point linear probe was employed to collect bulk conductivity measurements of the COF with
a space between tips of 1.25 mm. We calculated the bulk conductivity (S/cm) using Equation S1.
Herein, I (A) is current, V (V) is the voltage across the probes, s (cm) is distance between probes
(0.125 cm), F (unitless) is the correction factor accounting for the diameter and thickness of the
pellet.S14 The measurement gave a conductivity of 2.5110-5 S/cm.
𝐼 1
𝜎= (S1)
𝑉 2𝜋𝑠𝐹

Figure S21. (a) Representation of configuration for the measurement of conductivity by four-point
probe method and (b) the voltage (V) and current (A) recorded.

S22
To investigate the Arrhenius activation energy for electrical conductivity of COF-DC-8, a
2-point probe was employed to collect the current change under different temperature (293 K to
393 K) with a constant applied voltage of 1.0 V. Typically, the Arrhenius equation,
𝐸
− 𝑎
𝜎 = 𝜎0 𝑒 𝑘𝐵 𝑇 (S2)
can be used to access the Arrhenius activation energy.S15 In this equation, 𝜎 is the conductivity, σ0
is the pre-exponential factor and is constant, Ea is the Arrhenius activation energy, kB is the
Boltzmann constant, and T is the temperature. Because the current I measured is proportional to
the conductivity σ under the same temperature, the above equation can be further written as,
𝐸
− 𝑎
𝐼 = A𝑒 𝑘𝐵 𝑇
(S3),
which equals
1𝐸
𝑙𝑛(𝐼) = 𝑙𝑛(A) − 𝑇 𝑘𝑎 (S4)
𝐵

where A is a constant. The plot of natural log of current I (ln(I)) to reciprocal temperature (1/T, in
𝐸
K-1) generates the value − 𝑘𝑎 . The activation energy determined by this method was 324 meV.
𝐵

Figure S22. (a) The current change of a COF-DC-8 pellet measured by a 2-point probe with as the
temperature was increased from 293 to 393 K with an applied voltage of 1 V. (b) The Arrhenius
plot of the current and temperature.
For I2 doping studies, COF-DC-8 was drop casted onto a glass device equipped with
interdigitated gold electrodes. The devices was put into an I2 atmosphere (concentration ~400
ppm)S16 in a Teflon chamber under room temperature by using a similar method as previously
reported.S17 The current of the device was monitored continuously by a potentiostat under a constant
voltage bias of 1.0 V. The time-dependent study revealed that the current reached a maximal value
after exposure to I2 for 30 min with a 1000-fold increase. The doping process was very fast. 900-
fold increase in the current was observed within 5 min. The doping was partially reversible, as the

S23
current decreased to 36-fold of the initial value after the device was removed from the I2 chamber
and left under a continuous N2 flow.

Figure S23. The change in current of COF-DC-8 upon I2 doping.

12. Gas Sensing Experiments


12.1. Fabrication of Gas Sensors
1 mg of COF sample was dispersed in 1 mL of deionized water. The mixture was then
sonicated for 30 minutes and a suspension was obtained. 10 µL of the suspension was drop-cast
onto interdigitated gold electrodes with 5 μm gaps (part NO. G-IDEAU5, purchased form
Metrohm), which was then allowed to dry in air under room temperature in a dark area for about
18 hours before use. The resistances of the devices measured at room temperature under air using
a multimeter were 238±43 kΩ.

Figure S24. 8 chemiresistive devices made from COF-DC-8.

S24
12.2. Sensing Experiments
The experimental apparatus used in gas sensing is shown in the photograph (Figure S25a)
and schematic diagram (Figure S25b).

Figure S25. (a) Photograph and (b) schematic diagram of experimental apparatus used in gas
sensing experiments.

Generation of analytes with desired concentrations. To deliver the target gases with
concentration, a Sierra Micro-Trak and a Smart-Trak mass flow controller were used in
combination to deliver concentrations of gases from custom-ordered pre-mixed tanks (10,000 ppm
of NH3, H2S and NO in N2, 100 ppm of NO in N2, 10,000 ppm of NO2 in air) equipped with two-
stage stainless steel regulators. Gas streams from the tanks were diluted with dry N2 for delivery of
controlled concentrations of gases. The concentrations of target gases (NH3, NO, NO2 and H2S)
were adjusted by controlling the flow rates of the target gas and balance gas, which was achieved
using two mass-flow controllers. The flow of balance/purging gas was controlled at 0.5 or 1.0
L/min and the flow of the analyte was controlled at 0.2 to 4.0 mL/min. Before target gas exposure,
the gas sensors were stabilized under a constant flow of N2 for 30 min to get a flat base line. For

S25
the sensing test of NH3, NO, NO2, and H2S, the exposure of target gas and purging N2 gas were 30
minutes.

Current monitoring. The sensing performance of the fabricated sensors was monitored in a
sealed Teflon chamber at room temperature. An electrical feed-through and gas inlet and outlet
were installed in the chamber through which the analyte can pass through. The devices were
connected to the potentiostat through the bread board installed in the Teflon chamber. The current
through the device was monitored by a potentiostat by using the chronoamperometry analytical
method. The applied voltage was set at 1.0 V and interval for recording current was set at 0.5 s.

Data processing. Raw current data which shows the current change with time was
normalized and converted to change in normalized conductance according to Equation S5, wherein
Io = initial current and I = current at various points during measurement.

∆𝐺 𝐼−𝐼0
− =− × 100% (S5)
𝐺0 𝐼0

𝐼𝑒𝑛𝑑 −𝐼30
The reversibility of response is defined as , in which I30 is the current of the device
𝐼30 −𝐼0

after 30 min exposure to analytes and Iend is the current of the device after 30 min flushing of N2
after analyte exposure.

The theoretical limits of detection (LOD) were calculated using reported protocols.S18 The
noise-based deviation in −∆G/Go was calculated as the root mean squared (rms) value of the
baseline trace before exposure to analyte. Consecutive points (N ~ 1200) were selected and fitted
to a 5th order polynomial. The sum of squared residuals (SSR) 𝑉𝑥2 was then calculated using
Equation S6, where yi is measured −∆G/Go and y is the value calculated from the polynomial fit.
The root-mean-square deviation (rmsnoise) was calculated by Equation S7.

𝑉𝑥2 = ∑(𝑦𝑖 − 𝑦)2 (S6)

𝑉𝑥2
𝑟𝑚𝑠𝑛𝑜𝑖𝑠𝑒 = √ 𝑁
(S7)

The plots for concentration of analyte versus response (−∆G/Go) after a specific exposure
time were isolated wherein this relationship was linear (slope = m). The theoretical LOD was
calculated from Equation S8.

𝑟𝑚𝑠𝑛𝑜𝑖𝑠𝑒
𝐿𝑂𝐷 = 3 × 𝑚
(S8)

S26
12.2.1. NH3 Sensing

Figure S26. Responses of COF-DC-8 devices to (a) 80, (b) 40, (c) 20, (d) 10, (e) 5, and (f) 2 ppm
of NH3. For each concentration, 3–4 freshly prepared devices were used at an applied voltage of
1.0 V.

Figure S27. (a) Response of COF-DC-8 devices upon different exposure times as a function of
NH3 concentration at 2-80 ppm. The observed linear relationship between the response and
concentration upon (b) 1.5, (c) 5, (d) 10, (e) 20, and (f) 30 min exposure. SSR was 1.23377. N=1177,
RMS was 0.032376.

S27
Table S3. Reversibility of responses to 2-80 ppm of NH3.

concentration/ppm 2 5 10 20 40 80
Reversibility/% 97.8±2.1 63.4±7.2 79.9±2.1 66.0±7.2 52.9±6.2 60.6±13.3

Table S4. Calculated limits of detection for NH3 using COF-DC-8 as a function of exposure times.

exposure time/min 1.5 5 10 20 30


LOD/ppb 70.0 55.5 61.4 53.6 56.8

12.2.2. H2S Sensing

Figure S28. Responses of COF-DC-8 devices to (a) 80, (b) 40, (c) 20, (d) 10, (e) 5, and (f) 2 ppm
of H2S. For each concentration, 3–4 freshly prepared devices were used at an applied voltage of 1.0
V.

S28
Figure S29. (a) Response of COF-DC-8 devices upon different exposure times as a function of
H2S concentration at 2-80 ppm. The observed linear relationship between the response and
concentration upon (b) 1.5, (c) 5, (d) 10, (e) 20, and (f) 30 min exposure. SSR was 0.66777. N=1177,
RMS was 0.016244.

Table S5. Reversibility of responses to 2-80 ppm of H2S.

concentration/ppm 2 5 10 20 40 80
Reversibility/% -- 4.7±2.2 14.3±2.8 13±2.5 9.2±0.4 7.6±1.0

Table S6. Calculated limits of detection for H2S using COF-DC-8 as a function of exposure
times.

exposure time/min 1.5 5 10 20 30


LOD/ppb 204.2 155.3 140.1 127.6 121.1

S29
12.2.3. NO Sensing

Figure S30. Responses of COF-DC-8 devices to (a) 80, (b) 40, (c) 20, (d) 10, (e) 5, (f) 2 , (g) 1,
(h) 0.5, (i) 0.2, (j) 0.08, and (k) 0.02 ppm of NO. For each concentration, 3–4 freshly prepared
devices were used at an applied voltage of 1.0 V.

S30
Figure S31. (a) Response of COF-DC-8 devices upon different exposure time as a function of NO
concentration at 0.02-80 ppm. The observed linear relationship between the response and
concentration upon (b) 1.5, (c) 5, (d) 10, (e) 20, and (f) 30 min exposure. SSR was 1.57113. N=1201,
RMS was 0.036168.

Table S7. Reversibility of responses to 80-0.02 ppm of NO.

Concentration
0.02 0.08 0.2 0.5 1 2 5 10 20 40 80
/ppm
Reversibility 7.2± 15.8± 20.8± 22.0± 31.8± 19.7± 33.2± 28.0± 29.3± 26.8± 24.9±
/% 1.3 3.3 4.0 0.7 2.0 0.8 0.8 0.6 0.0 0.9 2.2

Table S8. Calculated limits of detection for NO using COF-DC-8 as a function of exposure
times.

exposure time/min 1.5 5 10 20 30


LOD/ppb 5.40 2.06 1.41 0.75 0.85

S31
12.2.4. NO2 Sensing

Figure S32. Responses of COF-DC-8 devices to (a) 80, (b) 40, (c) 20, (d) 10, (e) 5, and (f) 2 ppm
of NO2. For each concentration, 3–4 freshly prepared devices were used at an applied voltage of
1.0 V.

Figure S33. Response of COF-DC-8 device as a function of NO2 concentration. The response
values after the exposure for 1.5, 5, 10, 20, and 30 min were plotted. SSR was 1.30515. N=1170,
RMS was 0.0334.

S32
Table S9. Reversibility of responses to 2-80 ppm of NO2.

concentration/ppm 2 5 10 20 40 80
Reversibility/% 25.3±3.0 35.4±2.6 35.5±0.6 39.2±0.4 44.5±5.0 37.4±0.7

Table S10. Calculated limits of detection for NO2 using COF-DC-8 as a function of exposure times.

exposure time/min 1.5 5 10 20 30


LOD/ppb 15.93 4.03 1.86 0.88 0.61

Figure S34. Comparison of the response to (a) reducing gases NH3 and H2S and (b) oxidizing gases
NO and NO2 at 40 ppm.

12.2.5. Analysis of the Kinetic of Sensing Response


The rate and magnitude of the chemiresistive response towards analytes is a multifaced
function of at least five independent parameters, such as : (1) method of analyte delivery (constant
volume vs constant concentration);S19 (2) volume of sensing enclosure chamber;S20 (3) morphology,
preferential orientation, and film thickness of the material on the devices;S21 (4) contact area of the
materials with the device electrodes;S22 (5) the chemical nature of analyte-material interactions.S23
Based on the above hypothesis and analysis, the overall rate of the sensing of gas by COF-DC-8
is related to several factors, including the kinetics of the analyte-material interactions, the electronic
nature of the analyte, and the morphology of the material. In our study, the analyte delivery method,
the volume of the sensing chamber, and the sensing material were kept constant. As all the devices
were made by the same drop casting procedure, we also hypothesized that the variance of contact
area of the material with the device electrodes was negligible.

Typical sensing responses shown below (Figure S35) are usually composed of multiple
stages, including the initial rate and subsequent rates of response, in which the rate of response

S33
changes from fast to slow over time. At the initial stage of analyte exposure, the response changes
rapidly; with the increasing of dosing time, the increase in response progressively slows down, as
the sensor response approaches saturation (See Figure S26, S28 and S30a-e).

Figure S35. Typical sensing traces consisting of multiple stages showing that the response
increases at different rates over the course of an exposure event.

12.2.5.1. Initial Rate of Response


Analyzing the initial rate of response of the sensor is a convenient technique that allows
concentration-dependent measurement within a short timescale (less than a minute).S24 The site-
binding mechanism is usually used to the describe chemiresistive sensing, which hypothesizes that
the response is proportional to the concentration of the analyte bound by the sensing material.S25 At
the initial stage of an analyte exposure event, only a very small portion of the activate sites (for a
particular target analyte) on the surface of the materials are occupied, the interaction between the
analyte and active sites on the surface of the sensing material can be treated as a pseudo-first-order
reaction.S26 In this case, the initial rate of the analyte binding will be proportional to the
concentration of the analyte. Since the response of the device is proportional to the quantity of
analyte bound to the surface of the sensing material based on the site-binding mechanism,S25 the
initial rate of response is thus expected to be proportional to the concentration of the analyte near
the surface of the material. Shown in Figure S35-38 are the linear fits of the response of COF-DC-
8 to NH3, H2S, NO, and NO2 with 1 min exposure and the fitting of the relationship between the
initial rate of response and gas concentrations.

S34
Figure S36. (a) Responses of COF-DC-8 devices to 80, 40, 20, 10, 5, and 2 ppm of NH3 after 1
min exposure and the linear fitting of the response. (b) The slope of the fitting as a function of
concentration.

Figure S37. (a) Responses of COF-DC-8 devices to 80, 40, 20, 10, 5, and 2 ppm of H2S after 1
min exposure and the linear fitting of the response. (b) The slope of the fitting as a function of
concentration.

S35
Figure S38. (a) Responses of COF-DC-8 devices to 80-0.02 ppm of NO after 1 min exposure and
the linear fitting of the response. (b) The slope of the fitting as a function of concentration.

Figure S39. (a) Responses of COF-DC-8 devices to 80, 40, 20, 10, 5, and 2 ppm of NO2 after 1
min exposure and the linear fitting of the response. (b) The slope of the fitting as a function of
concentration.

We noticed that all four gases showed different initial rates of response, which followed
the orders of NH3>H2S (reducing gases) and NO>NO2 (oxidizing gases), meaning that specific
analyte-material interactions gave different initial rates of response. The efficiency of analyte-
material interaction is further affected by the kinetics of the analyte binding and charge transfer
efficiency between the analyte and sensing material.

12.2.5.2. Average Rate of Response


With most of the active sites directly exposed on the surface of the material occupied by
the analyte, the rate of analyte binding slowed down as available active sites decrease. For active
sites that are not directly exposed on the surface of the device, the analyte may need to diffuse

S36
across grain boundaries or travel through the interior pore to interact with the material. In these
cases, the diffusion rate of the analyte can affect the rate of the response as well.
Compared with the gas sensing performance of other conductive framework materials, we
found that the sensing dynamics of COF-DC-8 was comparable to the response dynamics of the
materials in the form of bulk powder,S12, S27 but slower than the sensing dynamic of thin film
devices.28 For sensing with device made of bulk form of material, analytes need to diffuse across
crystallite boundaries to interact with more active sites, which may slow down the average rate of
response. On the contrary, for devices made of thin films/sheets of sensing material, analyte can
rapidly diffuse across the surface of the material to interact with active sites which are substantially
exposed, thus triggering a rapid change in conductivity. This observation was also very recently
demonstrated by the Long group based on their study of chemiresisitivity in a conductive MOF, in
which the equilibration of the bulk composition appeared to have a longer time scale than the thin
pellet.S29
12.2.6. Recovery of the Sensing Performance by Thermal and Solvent Activation
The COF devices exhibited dosimetric responses toward H2S, NO and NO2. In order to
investigate the reusability of the devices, the used COF devices were reactivated by two means:
thermal recovery and solvent recovery. In thermal recovery, the devices that had been used after a
complete sensing test (N2 purge→30 min analyte exposure→30 min N2 purge) were heated at 60
ºC for 18 h under house vacuum (~20 mTorr). In solvent recovery, the devices were soaked in
deionized water for 1 hour, then allowed to dry overnight under air at room temperature.

Figure S40. Sensing response to (a) 80 ppm of H2S, (b) 40 ppm of NO, and (c) 40 ppm of NO2
with freshly made devices and devices after thermal and solvent activation.

After solvent recovery, device function was largely restored (108% for H2S, 71% for NO,
and 100% for NO2), while the devices only partially returned their performance after thermal
recovery (60% for H2S, 14% for NO, and 17% for NO2). These observations are consistent with

S37
the previous report that a washing procedure was more efficient than heating in the recovery of the
performance of Ni3HITP2 MOF coated fabric devices.27b During water treatment, the analyte
molecules bound to the material could be released released from the surface with the aid of water
or oxygen, which would restore the sensing ability of the COF-DC-8 device.

12.3. Sensing Mechanism by XPS and EPR


To investigate the sensing mechanism, XPS and electron paramagnetic resonance
spectroscopy were used. For both techniques, difference studies were utilized to observe changes
induced in the materials as a result of analyte exposure and binding. The two methods were used
to provide elemental confirmation of analyte binding (XPS) and analysis of oxidation state changes
within the framework (XPS and EPR).

Pristine materials were prepared for XPS analysis by placing 6 mg of COF powder into a
2 mL vial and degassing with N2 (1 L/min) for 1 h. Pristine samples were stored under N2. Analyte-
exposed materials were prepared by purging the pristine COF with N2 (1 L/min) for 1 hour followed
by analyte exposure (concentration of NH3, H2S, NO, and NO2 is 1%) for 1 hour. The samples were
then sealed under an atmosphere of 1% analyte (balanced with N2) overnight before testing. Powder
samples were mounted by lightly pressing onto copper tape which was mounted on a Dual-Height
(Kratos) sample holder. Each sample was first focused in the z-direction based on the C 1s emission
line. A survey spectrum was then obtained from 0 eV–1400 eV to obtain elemental surface
composition. High resolution spectra were then obtained at energy regions specific to elements
observed in the survey spectrum (C 1s, N 1s, O 1s, S 2p, Ni 2p 3/2). The XPS was performed under
a pressure of ~10−7 Torr.

To prepare EPR samples, ~2 mg of COF powder were transferred into a quartz EPR tube.
The quartz EPR tube was then degassed with N2 for (1 L/min) for 1 hour. After purging, the tube
was capped and transferred to the EPR instrument for analysis. The pristine samples were then
dosed with analyte to provide an analyte-exposed sample. These analyte-exposed samples were
prepared by a gas purging cycle of N2 (1 L/min, 1 hour)—analyte (1 hour)—N2 (1 L/min, 30
minutes) on pristine COF. The concentrations of the analytes are controlled at 1% for NH3, H2S,
NO, and NO2. EPR experiments were run in the X-band with microwave frequencies of 9.77 GHz.
Each experiment consisted of 8 scans across with the center field at 3500 G and sweep width of
3000 G. All experiments were run at 77 K using a square resonance chamber.

S38
Figure S41. Comparison of the XPS spectra of COF-DC-8 before and after NH3 dosing. (a)-(c)
The high resolution XPS spectra of C 1s, N 1s, and Ni 2p regions before NH3 dosing. (d)-(f) The
high resolution XPS spectra of C 1s, N 1s, and Ni 2p regions after NH3 dosing.

Figure S42. Comparison of the XPS spectra of COF-DC-8 before and after H2S dosing. (a)-(c)
The high resolution XPS spectra of C 1s, N 1s, and Ni 2p before H2S dosing. (d)-(g) The high
resolution XPS spectra of C 1s, N 1s, Ni 2p, and S 2p after H2S dosing.

S39
Figure S43. Comparison of the XPS spectra of COF-DC-8 before and after NO dosing. (a)-(c) The
high resolution XPS spectra of C 1s, N 1s, and Ni 2p before NO dosing. (d)-(f) The high resolution
XPS spectra of C 1s, N 1s, and Ni 2p after NO dosing.

Figure S44. Comparison of the XPS spectra of COF-DC-8 before and after NO2 dosing. (a)-(c)
The high resolution XPS spectra of C 1s, N 1s, and Ni 2p before NO2 dosing. (d)-(f) The high
resolution XPS spectra of C 1s, N 1s, and Ni 2p after NO2 dosing.

S40
Figure S45. Comparison of the EPR spectra of COF-DC-8 before (dashed lines) and after (solid
lines) dosing with (a) NH3, (b) H2S, (c) NO, and (d) NO2.

After dosing with NH3, no obvious change was detected in the XPS spectra of C 1s, N 1s,
and Ni 2p (Figure S41), meaning that the interaction between NH3 and DC-COF-8 may be
relatively weak and reversible under XPS experimental conditions. The EPR spectrum of the
pristine COF-DC-8 showed a ligand centered radical at g=1.99 (Figure S45), which is likely due
to the chemisorption of O2 on the Ni center of the phthalocyanine that induces unpaired radicals
because of the charge transfer from the phthalocyanine ligand to O2 molecule.S26, 30 After the
exposure of the NH3, only a slight intensity increase was observed (Figure S45a), indicating that
the interaction between NH3 and COF-DC-8 may be sufficiently reversible,S30b, 31 which was
consistent with XPS results.

In contrast, after dosing with H2S, clear peaks were observed at the S 2p binding energies.
The peaks of S 2p1/2 and S 2p3/2 at binding energies of 163.8 eV and 165.1 eV (Figure S42g),
similar with those found in metal sulfides,S32 indicate the formation of the Ni-S bond. This
assumption was also in accordance with the appearance of the satellite peak in the Ni 2p range
(Figure S42f), which may be induced by formation of Ni-S bond at the vertical direction. In the

S41
pristine COF-DC-8 the satellite peak was almost absent (Figure S42c), similar with those found
in small molecular Ni phthalocyanine and porphyrin.S7, 33 In addition, a prominent peak at higher
binding energy of 167.5 eV, which usually corresponds to sulphite species, was also observed
(Figure S42g). This observation suggested that the Ni bounded H2S may further transform into
species higher oxides, likely with the participation of surface bounded O2 through a surface-
promoted oxidation reaction.S32b In EPR experiments, we observed an obvious 70% intensity
decrease in the signal (Figure S45b), which indicated that H2S molecules may replace the O2
molecules and occupy the Ni centers.

After the exposure of the NO and NO2, no obvious changes were observed for spectra of Ni
2p (Figure S43f, S44f), indicating that the oxidation state of the Ni centers did not change during
the interaction between NO/NO2 and COF-DC-8. Deconvolution of the C 1s peak after NO dosing
showed only a slight increase in intensity at the binding energy of 287.5 eV (Figure S43d). A new
component was observed at the binding energy of 402.0 eV (Figure S43e), which was absent in
the pristine material. Based on the previous studies on the NO adsorption on metal surfaces, this
component can be assigned as a Ni-bound NO species.S34 While, the possibility that part of the NO
molecules may directly bind to the nickelphthalocyanine moiety cannot be excluded, as the N 1s
spectrum of NO adsorbed on organic surface may appear at a similar position.S35 After NO2 dosing,
a weak yet clear peak was observed at the binding energy of 406.1 eV (Figure S44e), suggesting
the existence of NO2 molecules which were likely coordinated to the Ni center. This
characterization was in accordance with literature values reported for NO2 bound to
metallophthalocyaninesS36 and metal oxides.S37 Prominent increases of 3.2 and 4.2 folds in the EPR
signals were found after NO and NO2 exposure (Figure S45c, S45d), respectively. These results
suggested ligand to ligand charge transfer interactions between the NO/NO2 and the phthalocyanine
component may occur upon analyte exposure.

13. References
S1. Hu, J.; Zhang, D.; Harris, F. W. J. Org. Chem. 2005, 70, 707.
S2. Shao, J.; Chang, J.; Chi, C. Linear and Star-Shaped Pyrazine-Containing Acene
Dicarboximides with High Electron-Affinity. Org. Biomol. Chem. 2012, 10, 7045.
S3. (a) Thelemann, J.; Illarionov, B.; Barylyuk, K.; Geist, J.; Kirchmair, J.; Schneider, P.;
Anthore, L.; Root, K.; Trapp, N.; Bacher, A.; Witschel, M.; Zenobi, R.; Fischer, M.; Schneider, G.;
Diederich, F. Aryl Bis-Sulfonamide Inhibitors of IspF from Arabidopsis Thaliana and Plasmodium
Falciparum. ChemMedChem 2015, 10, 2090; (b) Yüksel, F.; Gül Gürek, A.; Lebrun, C.; Ahsen, V.
Synthesis and Solvent Effects on the Spectroscopic Properties of Octatosylamido Phthalocyanines.
New J. Chem. 2005, 29, 726.
S4. Jia, H.; Yao, Y.; Zhao, J.; Gao, Y.; Luo, Z.; Du, P. A Novel Two-Dimensional Nickel
Phthalocyanine-Based Metal–Organic Framework for Highly Efficient Water Oxidation Catalysis.

S42
J. Mater. Chem. A 2018, 6, 1188.
S5. Lozzi, L.; Ottaviano, L.; Santucci, S. High Resolution XPS Studies on Hexadecafluoro-
Copper-Phthalocyanine Deposited onto Si(111)7×7 Surface. Surf. Sci. 2001, 470, 265.
S6. Mahmood, J.; Lee, E. K.; Jung, M.; Shin, D.; Jeon, I. Y.; Jung, S. M.; Choi, H. J.; Seo, J.
M.; Bae, S. Y.; Sohn, S. D.; Park, N.; Oh, J. H.; Shin, H. J.; Baek, J. B. Nitrogenated Holey Two-
Dimensional Structures. Nat. Commun. 2015, 6, 6486.
S7. Ottaviano, L.; Di Nardo, S.; Lozzi, L.; Passacantando, M.; Picozzi, P.; Santucci, S. Thin
and Ultra-Thin Films of Nickel Phthalocyanine Grown on Highly Oriented Pyrolitic Graphite: An
XPS, UHV-AFM and Air Tapping-Mode AFM Study. Surf. Sci. 1997, 373, 318.
S8. Clough, A. J.; Skelton, J. M.; Downes, C. A.; de la Rosa, A. A.; Yoo, J. W.; Walsh, A.;
Melot, B. C.; Marinescu, S. C. Metallic Conductivity in a Two-Dimensional Cobalt Dithiolene
Metal-Organic Framework. J. Am. Chem. Soc. 2017, 139, 10863.
S9. (a) Ding, X.; Feng, X.; Saeki, A.; Seki, S.; Nagai, A.; Jiang, D. Conducting
Metallophthalocyanine 2d Covalent Organic Frameworks: The Role of Central Metals in
Controlling π-Electronic Functions. Chem. Commun. 2012, 48, 8952; (b) Jin, S.; Ding, X.; Feng,
X.; Supur, M.; Furukawa, K.; Takahashi, S.; Addicoat, M.; El-Khouly, M. E.; Nakamura, T.; Irle,
S.; Fukuzumi, S.; Nagai, A.; Jiang, D. Charge Dynamics in a Donor-Acceptor Covalent Organic
Framework with Periodically Ordered Bicontinuous Heterojunctions. Angew. Chem. Int. Ed. 2013,
52, 2017.
S10. (a) Wan, S.; Gándara, F.; Asano, A.; Furukawa, H.; Saeki, A.; Dey, S. K.; Liao, L.;
Ambrogio, M. W.; Botros, Y. Y.; Duan, X.; Seki, S.; Stoddart, J. F.; Yaghi, O. M. Covalent Organic
Frameworks with High Charge Carrier Mobility. Chem. Mater. 2011, 23, 4094; (b) Feng, X.; Liu,
L.; Honsho, Y.; Saeki, A.; Seki, S.; Irle, S.; Dong, Y.; Nagai, A.; Jiang, D. High-Rate Charge-Carrier
Transport in Porphyrin Covalent Organic Frameworks: Switching from Hole to Electron to
Ambipolar Conduction. Angew. Chem. Int. Ed. 2012, 51, 2618.
S11. (a) Wiktor, C.; Meledina, M.; Turner, S.; Lebedev, O. I.; Fischer, R. A. Transmission
Electron Microscopy on Metal–Organic Frameworks – a Review. J. Mater. Chem. A 2017, 5, 14969;
(b) Hmadeh, M.; Lu, Z.; Liu, Z.; Gándara, F.; Furukawa, H.; Wan, S.; Augustyn, V.; Chang, R.;
Liao, L.; Zhou, F.; Perre, E.; Ozolins, V.; Suenaga, K.; Duan, X.; Dunn, B.; Yamamto, Y.; Terasaki,
O.; Yaghi, O. M. New Porous Crystals of Extended Metal-Catecholates. Chem. Mater. 2012, 24,
3511; (c) Zhang, D.; Zhu, Y.; Liu, L.; Ying, X.; Hsiung, C. E.; Sougrat, R.; Li, K.; Han, Y. Atomic-
Resolution Transmission Electron Microscopy of Electron Beam-Sensitive Crystalline Materials.
Science 2018, 359, 675.
S12. Meng, Z.; Aykanat, A.; Mirica, K. A. Welding Metallophthalocyanines into Bimetallic
Molecular Meshes for Ultrasensitive, Low-Power Chemiresistive Detection of Gases. J. Am. Chem.
Soc. 2019, 141, 2046.
S13. Nagatomi, H.; Yanai, N.; Yamada, T.; Shiraishi, K.; Kimizuka, N. Synthesis and Electric
Properties of a Two-Dimensional Metal-Organic Framework Based on Phthalocyanine. Chem. Eur.
J. 2018, 24, 1806.
S14. Schroder, D. K. Semiconductor Material and Device Characterization. John Wiley & Sons,
Inc.: 2005.
S15. Darago, L. E.; Aubrey, M. L.; Yu, C. J.; Gonzalez, M. I.; Long, J. R. Electronic Conductivity,
Ferrimagnetic Ordering, and Reductive Insertion Mediated by Organic Mixed-Valence in a Ferric
Semiquinoid Metal-Organic Framework. J. Am. Chem. Soc. 2015, 137, 15703.
S16. Baxter, G. P.; Hickey, C. H.; Holmes, W. C. The Vapor Pressure of Iodine. J. Am. Chem.
Soc. 1907, 29, 127.
S17. (a) Wan, S.; Guo, J.; Kim, J.; Ihee, H.; Jiang, D. A Belt-Shaped, Blue Luminescent, and
Semiconducting Covalent Organic Framework. Angew. Chem. Int. Ed. 2008, 47, 8826; (b) Cai, S.-
L.; Zhang, Y.-B.; Pun, A. B.; He, B.; Yang, J.; Toma, F. M.; Sharp, I. D.; Yaghi, O. M.; Fan, J.;
Zheng, S.-R.; Zhang, W.-G.; Liu, Y. Tunable Electrical Conductivity in Oriented Thin Films of
Tetrathiafulvalene-Based Covalent Organic Framework. Chem. Sci. 2014, 5, 4693.

S43
S18. (a) Meng, Z.; Aykanat, A.; Mirica, K. A. Welding Metallophthalocyanines into Bimetallic
Molecular Meshes for Ultrasensitive, Low-Power Chemiresistive Detection of Gases. J. Am. Chem.
Soc. 2019, 141, 2046; (b) Li, J.; Lu, Y.; Ye, Q.; Cinke, M.; Han, J.; Meyyappan, M. Carbon
Nanotube Sensors for Gas and Organic Vapor Detection. Nano Lett. 2003, 3, 929; (c) Ammu, S.;
Dua, V.; Agnihotra, S. R.; Surwade, S. P.; Phulgirkar, A.; Patel, S.; Manohar, S. K. Flexible, All-
Organic Chemiresistor for Detecting Chemically Aggressive Vapors. J. Am. Chem. Soc. 2012, 134,
4553.
S19. (a) Pandey, S. Highly Sensitive and Selective Chemiresistor Gas/Vapor Sensors Based on
Polyaniline Nanocomposite: A Comprehensive Review. J. Sci. Adv. Mater. Dev. 2016, 1, 431; (b)
Bertoni, C.; Naclerio, P.; Viviani, E.; Dal Zilio, S.; Carrato, S.; Fraleoni-Morgera, A.
Nanostructured P3HT as a Promising Sensingelement for Real-Time, Dynamic Detection
Ofgaseous Acetone. Sensors 2019, 19, 1296.
S20. Reghu, A.; LeGore, L. J.; Vetelino, J. F.; Lad, R. J.; Frederick, B. G. Distinguishing Bulk
Conduction from Band Bending Transduction Mechanisms in Chemiresistive Metal Oxide Gas
Sensors. J. Phys. Chem. C 2018, 122, 10607.
S21. (a) Hoppe, B.; Hindricks, K. D. J.; Warwas, D. P.; Schulze, H. A.; Mohmeyer, A.; Pinkvos,
T. J.; Zailskas, S.; Krey, M. R.; Belke, C.; König, S.; Fröba, M.; Haug, R. J.; Behrens, P. Graphene-
Like Metal–Organic Frameworks: Morphology Control, Optimization of Thin Film Electrical
Conductivity and Fast Sensing Applications. CrystEngComm 2018, 20, 6458; (b) Gargiulo, V.;
Alfano, B.; Di Capua, R.; Alfé, M.; Vorokhta, M.; Polichetti, T.; Massera, E.; Miglietta, M. L.;
Schiattarella, C.; Di Francia, G. Graphene-Like Layers as Promising Chemiresistive Sensing
Material for Detection of Alcohols at Low Concentration. J. Appl. Phys. 2018, 123, 024503; (c)
Donarelli, M.; Ottaviano, L.; Giancaterini, L.; Fioravanti, G.; Perrozzi, F.; Cantalini, C. Exfoliated
Black Phosphorus Gas Sensing Properties at Room Temperature. 2D Mater. 2016, 3, 025002.
S22. (a) Barsan, N.; Weimar, U. Understanding the Fundamental Principles of Metal Oxide
Based Gas Sensors; the Example of CO Sensing with SnO2 Sensors in the Presence of Humidity. J
Phys-Condens Mat. 2003, 15, R813; (b) Zhang, P. Design and Fabrication of Chemiresistor
Typemicro/Nano Hydrogen Gas Sensors Using Interdigitated Electrodes. Electronic Theses and
Dissertations 2008, 3705.
S23. Meng, Z.; Stolz, R. M.; Mendecki, L.; Mirica, K. A. Electrically-Transduced Chemical
Sensors Based on Two-Dimensional Nanomaterials. Chem. Rev. 2019, 119, 478.
S24. (a) Riedel, K.; Renneberg, R.; Khn, M.; Scheller, F. A Fast Estimation of Biochemical
Oxygen Demand Using Microbial Sensors. Appl. Microbiol. Biotechnol. 1988, 28; (b) Tan, T. C.;
Li, F.; Neoh, K. G. Measurement of BOD by Initial Rate of Response of a Microbial Sensor. Sensors
Actuators B: Chem. 1993, 10, 137.
S25. Kong, J.; Javey, A. Carbon Nanotube Electronics. Springer, 2009.
S26. Bohrer, F. I.; Sharoni, A.; Colesniuc, C.; Park, J.; Schuller, I. K.; Kummel, A. C.; Trogler,
W. C. Gas Sensing Mechanism in Chemiresistive Cobalt and Metal-Free Phthalocyanine Thin
Films. J. Am. Chem. Soc. 2007, 129, 5640.
S27. (a) Smith, M. K.; Jensen, K. E.; Pivak, P. A.; Mirica, K. A. Direct Self-Assembly of
Conductive Nanorods of Metal–Organic Frameworks into Chemiresistive Devices on Shrinkable
Polymer Films. Chem. Mater. 2016, 28, 5264; (b) Smith, M. K.; Mirica, K. A. Self-Organized
Frameworks on Textiles (SOFT): Conductive Fabrics for Simultaneous Sensing, Capture, and
Filtration of Gases. J. Am. Chem. Soc. 2017, 139, 16759.
S28. (a) Yao, M.-S.; Lv, X.-J.; Fu, Z.-H.; Li, W.-H.; Deng, W.-H.; Wu, G.-D.; Xu, G. Layer-by-
Layer Assembled Conductive Metal-Organic Framework Nanofilms for Room-Temperature
Chemiresistive Sensing. Angew. Chem. Int. Ed. 2017, 129, 16737; (b) Li, S. C.; Lu, T. Y.; Zheng, J.
C.; Yang, S. W.; Wang, J. S.; Wu, G. Origin of Metallicity in 2D Multilayer Nickel Bis(Dithiolene)
Sheets. 2D Mater. 2018, 5, 035027.
S29. Aubrey, M. L.; Kapelewski, M. T.; Melville, J. F.; Oktawiec, J.; Presti, D.; Gagliardi, L.;
Long, J. R. Chemiresistive Detection of Gaseous Hydrocarbons and Interrogation of Charge

S44
Transport in Cu[Ni(2,3-Pyrazinedithiolate)2] by Gas Adsorption. J. Am. Chem. Soc. 2019, 141,
5005.
S30. (a) Bohrer, F. I.; Colesniuc, C. N.; Park, J.; Ruidiaz, M. E.; Schuller, I. K.; Kummel, A. C.;
Trogler, W. C. Comparative Gas Sensing in Cobalt, Nickel, Copper, Zinc, and Metal-Free
Phthalocyanine Chemiresistors. J. Am. Chem. Soc. 2009, 131, 478; (b) Collins, R. A.; Mohammed,
K. A. Gas Sensitivity of Some Metal Phthalocyanines. J. Phys. D: Appl. Phys. 1988, 21, 154.
S31. (a) Zou, D.; Zhao, W.; Cui, B.; Li, D.; Liu, D. Adsorption of Gas Molecules on a Manganese
Phthalocyanine Molecular Device and Its Possibility as a Gas Sensor. Phys. Chem. Chem. Phys.
2018, 20, 2048; (b) Klyamer, D. D.; Sukhikh, A. S.; Krasnov, P. O.; Gromilov, S. A.; Morozova, N.
B.; Basova, T. V. Thin Films of Tetrafluorosubstituted Cobalt Phthalocyanine: Structure and Sensor
Properties. Appl. Surf. Sci. 2016, 372, 79.
S32. (a) Ko, T. H.; Chu, H. Spectroscopic Study on Sorption of Hydrogen Sulfide by Means of
Red Soil. Spectrochim. Acta. A Mol. Biomol. Spectrosc. 2005, 61, 2253; (b) Galtayries, A.; Bonnelle,
J. P. XPS and ISS Studies on the Interaction of H2S with Polycrystalline Cu, Cu2O and CuO
Surfaces. Surf. Interface Anal. 1995, 23, 171; (c) Zhang, X.; Dou, G. Y.; Wang, Z.; Li, L.; Wang, Y.
F.; Wang, H. L.; Hao, Z. P. Selective Catalytic Oxidation of H2S over Iron Oxide Supported on
Alumina-Intercalated Laponite Clay Catalysts. J. Hazard. Mater. 2013, 260, 104.
S33. Berrios, C.; Marco, J. F.; Gutierrez, C.; Ureta-Zanartu, M. S. Study by XPS and Uv-Visible
and Drift Spectroscopies of Electropolymerized Films of Substituted Ni(II)-p-Phenylporphyrins
and -Phthalocyanines. J. Phys. Chem. B 2008, 112, 12644.
S34. (a) Herranz, T.; Deng, X.; Cabot, A.; Liu, Z.; Salmeron, M. In Situ XPS Study of the
Adsorption and Reactions of NO and O2 on Gold Nanoparticles Deposited on TiO2 and SiO2. J.
Catal. 2011, 283, 119; (b) Venezia, A. M.; Liotta, L. F.; Deganello, G.; Terreros, P.; Peña, M. A.;
Fierro, J. L. G. IR and XPS Study of NO and CO Interaction with Palladium Catalysts Supported
on Aluminosilicates. Langmuir 1999, 15, 1176; (c) Bukhtiyarov, A. V.; Kvon, R. I.; Nartova, A. V.;
Bukhtiyarov, V. I. An XPS and STM Study of the Size Effect in NO Adsorption on Gold
Nanoparticles. Russ. Chem. Bull. 2012, 60, 1977.
S35. Sysoev, V. I.; Okotrub, A. V.; Gusel'nikov, A. V.; Smirnov, D. A.; Bulusheva, L. G. In Situ
XPS Observation of Selective NOx Adsorption on the Oxygenated Graphene Films. Phys. Status.
Solidi. B 2018, 255, 1700267.
S36. (a) Mrwa, A.; Friedrich, M.; Hofmann, A.; Zahn, D. R. T. Response of Lead Phthalocyanine
to High NO2 Concentration. Sensors Actuators B: Chem. 1995, 25, 596; (b) Mockert, H.; Graf, K.;
Schmeisser, D.; Göpel, W.; Ahmad, Z. A.; Archer, P. B. M.; Chadwick, A. V.; Wright, J. D.
Characterization of Gas-Sensitive Lead Phthalocyanine Film Surfaces by X-Ray Photoelectron
Spectroscopy. Sensors Actuators B: Chem. 1990, 2, 133.
S37. Baltrusaitis, J.; Jayaweera, P. M.; Grassian, V. H. XPS Study of Nitrogen Dioxide
Adsorption on Metal Oxide Particle Surfaces under Different Environmental Conditions. Phys.
Chem. Chem. Phys. 2009, 11, 8295.

S45

You might also like