You are on page 1of 8

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: Org. Process Res. Dev. XXXX, XXX, XXX−XXX pubs.acs.org/OPRD

Explosion Hazards of Sodium Hydride in Dimethyl Sulfoxide,


N,N‑Dimethylformamide, and N,N‑Dimethylacetamide
Qiang Yang,*,† Min Sheng,‡ James J. Henkelis,† Siyu Tu,† Eric Wiensch,† Honglu Zhang,†
Yiqun Zhang,§ Craig Tucker,‡ and David E. Ejeh‡

Product Design & Process R&D, Corteva Agriscience, 9330 Zionsville Road, Indianapolis, Indiana 46268, United States

Reactive Chemicals, Corteva Agriscience, Midland, Michigan 48667, United States
§
Analytical Sciences, The Dow Chemical Company, Midland, Michigan 48667, United States
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The hazards associated with the thermal decomposition of chemically incompatible sodium hydride solvent
matrices are known, with reports from the 1960s detailing the inherent instability of NaH/dimethyl sulfoxide, NaH/N,N-
dimethylformamide, and NaH/N,N-dimethylacetamide mixtures. However, these hazards remain underappreciated and
undercommunicated, likely as a consequence of the widespread use of these NaH/solvent matrices in synthetic chemistry. We
Downloaded via 94.231.219.180 on August 10, 2019 at 13:26:58 (UTC).

report herein detailed investigations into the thermal stability of these mixtures and studies of the formation of gaseous products
from their thermal decomposition. We expect this contribution to promote awareness of these hazards within the wider
scientific community, encourage scientists to identify and pursue safer alternatives, and most importantly, help to prevent
incidents associated with these reactive mixtures.
KEYWORDS: sodium hydride (NaH), dimethyl sulfoxide (DMSO), N,N-dimethylformamide (DMF),
N,N-dimethylacetamide (DMAc), thermal instability, explosion

■ INTRODUCTION
The use of NaH as a strong base is commonplace in synthetic
maintained at 70 °C for 1 h and then cooled, during which
time they observed a sharp temperature spike followed by an
organic chemistry.1 Generally speaking, NaH is easy to handle explosion that was accompanied by a noxious gas. Immediately
after the disclosure of this incident, an analogous explosion
and store and often offers predictable reactivity and high atom
from a NaH/DMSO mixture was reported,8 albeit at a lower
efficiency. As a result, it has been used to facilitate a wide range
temperature (50 °C) and concentration (0.17:1 NaH/DMSO
of chemical transformations.2 The most commonly used
on a molar basis) compared with the conditions employed in
solvents for reactions that use NaH are dimethyl sulfoxide
the earlier incident.6
(DMSO), N,N-dimethylformamide (DMF), and N,N-dime-
The thermal decomposition of a NaH/DMF mixture was
thylacetamide (DMAc) because of their complementary
first observed in the early 1960s, first by Powers et al., who
physical properties and their ability to solubilize both organic
noted energetic decomposition during a NaH-mediated
and inorganic compounds. cleavage of a formyl group,9 and Brimacombe et al. during
The hazards associated with the thermal instability of NaH/ the alkylation of a carbohydrate using NaH as the base.10
DMSO, NaH/DMF, and NaH/DMAc mixtures have been a Nasipuri et al. later reported the aromatization of various
growing concern within the chemistry community.3 In addition cyclohexenones using NaH in DMF, in which the authors
to the desirable physical properties of DMSO to solubilize acknowledged the decomposition of DMF in the presence of
both organic and inorganic compounds, dimsyl ion (also NaH but still carried out their reactions at 100 °C.11 The
known as sodium dimsylate or methyl sulfinyl carbanion), researchers proposed that the active base in these reactions was
prepared via the reaction of NaH with DMSO under nitrogen sodium dimethylamine, resulting from the reaction of DMF
at 70−75 °C, has found wide applications in a variety of with NaH.
chemical transformations since the seminal publication by The hazards associated with the thermal instability of NaH/
Corey and Chaykovsky in the early 1960s.4 The thermal DMF pose an even greater risk as the reaction scale increases.
instability of dimsyl ion was first highlighted by Corey and Reports of large-scale incidents associated with the thermal
Chaykovsky in 19654b and further studied by many research instability of NaH/DMF were noted in the early 1980s. In the
groups.5 The first reported explosion resulting from a NaH/ first reported example, conducted on a pilot-plant scale at
DMSO mixture was in 1966,6 when scientists at the Cancer SmithKline Beckman, a condensation reaction employing NaH
Chemotherapy Research Department of Mount Zion Hospital in DMF experienced uncontrolled self-heating on 2500 L scale
and Medical Center (Palo Alto, CA) were scaling up a that resulted in the failure of a reactor’s rupture disk.12
methylation reaction that employed NaH in DMSO, following
a method reported by Russell and Weiner.7 The scientists
Special Issue: Corteva Agriscience
noted incomplete conversion when the reaction was carried
out using a NaH/DMSO molar ratio of 0.19:1 and thus Received: June 18, 2019
increased the ratio to 0.24:1. The reaction mixture was

© XXXX American Chemical Society A DOI: 10.1021/acs.oprd.9b00276


Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Article

Through a series of calorimetric studies, the research team


observed that mixtures of NaH with either DMF or DMAc
would begin to self-heat at temperatures as low as 26 °C and
that the temperature would rapidly increase to 80 °C, at which
point the cooling capacity of the pilot-plant reactor would be
exceeded. They further commented that the onset temperature
of decomposition was between 40 and 50 °C with carefully
dried DMF; however, the detailed thermal stability data
resulting from the study were not reported. A second account
was reported by Burdick and Jackson Laboratories in the same
year, stating that uncontrolled heating of a NaH/DMF reaction
mixture began at 40 °C and that the reaction temperature
quickly rose to more than 100 °C in less than 10 min.13
Although full cooling was applied to the reaction vessel, the
temperature could not be controlled, and the majority of the
DMF solvent evaporated before the reaction subsided. They Figure 1. Numbers of publications using NaH/DMSO, NaH/DMF,
commented that the reaction had been conducted many times and NaH/DMAc in Organic Letters in 2014−2018.
without an incident, which highlights the unpredictable nature
of this chemistry. While they suspected the reactor’s material of
construction (stainless steel) to be the likely cause of the
decomposition, this hypothesis was proven invalid by later
studies and data presented in this contribution.
The safety concerns regarding the thermal instability of NaH
mixtures were reiterated in two editorials by the previous
Organic Process Research & Development Editor-in-Chief,
Trevor Laird,14 who commented that multiple runaway
reactions with NaH/DMF mixtures had occurred before the
first report was published in 1982. These editorials drew much-
needed attention to NaH-containing reactive mixtures and
highlighted the importance of scientific communication to
raise awareness of these hazards and alert others to such
dangers.
Despite the numerous reports intended to caution scientists,
the hazards associated with NaH/DMSO, NaH/DMF, and
NaH/DMAc reactive mixtures remain underappreciated and Figure 2. Numbers of publications using NaH/DMSO, NaH/DMF,
and NaH/DMAc in The Journal of Organic Chemistry in 2014−2018.
inadequately discussed, overshadowed by the sheer number of
publications that utilize these hazardous conditions, further
diminishing their impact upon the chemistry community. As
shown in Figures 1−3, the number of publications reporting
the use of these hazardous conditions each year remains
consistently high. Between 2014 and 2018, Organic Letters
published 38−62 examples/year (Figure 1), whereas The
Journal of Organic Chemistry (Figure 2) and the Journal of
Medicinal Chemistry (Figure 3) published 28−46 and 67−94
examples/year, respectively.
It should be noted that most of the papers published in
Organic Process Research & Development employing NaH/
DMSO, NaH/DMF, and NaH/DMAc only reported these
hazardous conditions in their original discovery-phase routes
(Figure 4). In most instances, the authors acknowledged the
inherent safety hazards associated with these reactive mixtures
and designed their processes away from these mixtures in the
final routes. For example, Dahl et al.15 employed sodium Figure 3. Numbers of publications using NaH/DMSO, NaH/DMF,
dimsylate as the base in their process for the synthesis of (S)-2- and NaH/DMAc in the Journal of Medicinal Chemistry in 2014−2018.
({3-[(S)-5-chloro-1-(4-chlorophenyl)indan-1-yl]propyl}-
methylamino)propionic acid. The researchers performed a
detailed reaction calorimetry evaluation, which indicated that developed a safer process in which the reaction was performed
the sodium dimsylate decomposed at temperatures as low as using 4 equiv of DMSO (relative to NaH) in 7 volumes of
∼50 °C with an adiabatic temperature rise (ΔTad) of 500 °C in THF (relative to DMSO) and successfully scaled up this
pure DMSO. In contrast, when the DMSO solution of sodium process to 4.1 kg of 60 wt % NaH in mineral oil. Nonetheless,
dimsylate was diluted with THF, the lowest observed there are examples where these hazardous mixtures were
decomposition temperature was ∼100 °C with a ΔTad of employed in the final reaction conditions, one of which
230 °C. On the basis of these results, the researchers includes a large-scale production that operated using 880 mol
B DOI: 10.1021/acs.oprd.9b00276
Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development


Article

RESULTS AND DISCUSSION


An exothermic reaction is typically maintained at the optimal
temperature by applying external cooling until reaction
completion and then cooled for workup and isolation of the
product. However, in cases of insufficient cooling or cooling
failure, unreacted components will further react to release
energy, leading to a higher reaction temperature. In a worst-
case scenario (e.g., all reactants are accumulated and the
reaction continues under adiabatic conditions), the reaction
temperature would reach the maximum temperature of a
synthesis reaction (MTSR), which is equal to the desired
reaction temperature plus ΔTad.17 In cases when the MTSR
exceeds the onset temperature of reaction mixture decom-
position, a secondary reaction would be initiated, potentially
leading to runaway scenarios. It is thus crucial to evaluate
Figure 4. Numbers of publications using NaH/DMSO, NaH/DMF, reaction calorimetry and thermal stability hazards of chemical
and NaH/DMAc in Organic Process Research & Development in 1997− reactions and to design control strategies accordingly in order
2018. to ensure that thermal decomposition of the reaction mixture is
not triggered. Measurements of the heat of the desired reaction
and ΔTad are generally performed using an instrument such as
an RC1 reaction calorimeter, a micro reaction calorimeter
of NaH (35.2 kg of 60 wt % NaH in mineral oil).16 In these (μRC), or an EasyMax HF Cal heat flow calorimeter. The
cases the authors restated that alternative reaction conditions thermal stability of reaction mixtures is generally characterized
were not appropriate and that all of the reactions were with a differential scanning calorimetry (DSC), a thermal
conducted within safe operating ranges and in a reactor made screening unit (Tsu), an accelerating rate calorimetry (ARC),
of compatible materials of construction.8 It is worth noting that and/or a vent-sizing package (VSP).
calorimetry studies have proven that these hazards are 1. Thermal Stability Evaluation of NaH/DMSO. ARC
independent of the reactor material of construction (stainless analysis of 4.55 g of a mixture of 9.7% NaH, 6.4% mineral oil,
steel). Furthermore, critical reaction parameters, such as and 83.9% DMSO18 recorded two small exothermic events
residual water, reaction temperature, and relative concentration followed by a significant exothermic event with an onset
of NaH, might help mitigate some of the hazards associated temperature of 56.8 °C (Figure 5 and Table 1). This exotherm
with their use, but they do not completely address the dangers caused the rupture of an ARC cell designed with an average
of spontaneous thermal decomposition. burst pressure of 14 500 psi (Figure 6, left). The force
This literature survey shows that reactions with these generated from this explosion was strong enough to displace
matrices continue to appear in publications, indicating their the reactor housing, which sat on top of the ARC reactor
(Figure 6, right). The full heat and pressure information on
relatively common use and suggesting that the safety hazards
these exothermic events could not be recorded because of this
associated with the thermal instability of NaH/DMSO, NaH/ rupture.
DMF, and NaH/DMAc mixtures remain underappreciated. In A mixture of 10.3% NaH, 6.9% mineral oil, and 82.8%
this contribution, we communicate the results of detailed DMSO was then studied by DSC to better understand the
thermal stability evaluations of these hazardous combinations total heat output that could result from an exothermic event
in order to promote awareness of these hazards within the strong enough to rupture an ARC cell. As shown in Figure 7,
chemistry community and encourage scientists to employ safer DSC analysis recorded an endothermic event (14 to 39 °C), a
reaction conditions, with the ultimate goal of preventing minor exothermic event (39 to 120 °C), and two consecutive
incidents relating to these hazards. major exothermic events (120 to 285 °C). The first major

Figure 5. ARC heat rate and pressure vs temperature profiles of a NaH/DMSO mixture.

C DOI: 10.1021/acs.oprd.9b00276
Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Article

Table 1. ARC Analysis of a NaH/DMSO Mixture


sample description 9.7% NaH + 6.4% mineral oil + 83.9%
DMSO
total sample mass (g) 4.5471
Cp of sample (J g−1 °C−1) 1.923
ARC cell mass (g) 21.9702
set end temperature (°C) 350
phi 2.06
onset temperature (°C) 56.8
peak temperature (°C) −
end temperature (°C) −
max self-heating rate −
(°C/min)
total heat (J/g) ruptured Figure 8. ARC temperature and pressure vs time profiles of a NaH/
DMSO mixture.

overlaid in red (air blank) and black (sample of interest), are


compared in Figure 9. The gases detected in the sample

Figure 6. Pictures of (left) a ruptured Hastelloy C ARC cell and


(right) the displaced ARC reactor housing resulting from the cell
explosion.

Figure 9. Headspace GC/MS analysis of the gaseous products from


the thermal decomposition of a NaH/DMSO mixture.

referenced against the NIST chemical library supported the


presence of ethylene and dimethyl sulfide as the decomposition
products,20 along with DMSO.21 Undoubtedly, the generation
of large volumes of gases poses a significant safety concern and
contributed to the explosion during the dominant exothermic
Figure 7. DSC analysis of a NaH/DMSO mixture. event.
2. Thermal Stability Evaluation of NaH/DMF. The
exothermic event had a peak temperature of 146 °C and a total thermal stability of NaH/DMF mixtures was evaluated by ARC
heat of −413.5 J/g, while the second major exothermic event analysis. As shown in Figure 10, a significant exothermic
had a peak at 255 °C and a total heat of −810.2 J/g. With the decomposition with an onset temperature of 76.1 °C was
combined heat of these two events (−1123.7 J/g), this mixture detected with a mixture of 9.5% NaH, 6.3% mineral oil, and
was considered to be explosive according to the Yoshida 84.2% DMF. This decomposition produced an exotherm of
correlation.19 As a direct comparison, pure DMSO decom- −528.4 J/g with a maximum self-heating rate of 7.23 °C/min.
poses at around 278 °C with a heat output of −911.59 J/g by Increasing the weight ratio of NaH to 24.6% (in 16.4% mineral
DSC analysis. oil and 59% DMF) significantly lowered the decomposition
While the first of the two small exothermic events recorded onset temperature to 39.8 °C and increased the peak self-
by ARC (21.3 to 26.4 °C, EXO1 in Figure 8) was attributed to heating rate to 634.7 °C/min, affording a total exotherm of
the heat of mixing, the second one (36.9 to 43.5 °C, EXO2 in greater than −601.8 J/g (Table 2). Both tests had substantial
Figure 8), with a total heat of only −93.2 J/g, was most likely cool-down pressures, confirming that the decomposition of
caused by the reaction between NaH and DMSO. Surprisingly, DMF resulted in the formation of significant amounts of
this minor exotherm generated a significant amount of gaseous products.
noncondensable gases that raised the cell pressure from 129 Identification of the gaseous products resulting from the
to 1289 psi between the end of the first exothermic event (109 thermal decomposition of NaH/DMF was performed using
min) and the start of the major exothermic event (635 min, evolved gas analysis (EGA) Micro-GC analysis. In a control
EXO3 in Figure 8). experiment with NaH in mineral oil, only the residual water in
Analysis of the gaseous products from EXO2 was performed the EGA Micro-GC system, transfer lines, and sample cup was
using headspace GC/MS. The total ion chromatograms (TIC), detected at <100 °C. Release of H2 gas was detected when the
D DOI: 10.1021/acs.oprd.9b00276
Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Article

Figure 10. ARC heat rate and pressure vs temperature profiles of NaH/DMF mixtures.

Table 2. ARC Analysis of NaH/DMF Mixtures


sample description 9.5% NaH + 6.3% mineral oil + 84.2% DMF 24.6% NaH + 16.4% mineral oil + 59% DMF
total sample mass (g) 4.1234 3.3896
Cp of sample (J g−1 °C−1) 2.010 1.930
ARC cell mass (g) 21.8926 14.8079
set end temperature (°C) 350 200
phi 2.11 1.95
onset temperature (°C) 76.1 39.8
peak temperature (°C) 133.8 126.2
end temperature (°C) 200.7 >199.7
max self-heating rate (°C/min) 7.23 634.7
total heat output (J/g) −528.4 >−601.8

temperature reached ∼165 °C and ended at ∼360 °C,


indicating that the formation of H2 was caused by the thermal
decomposition of NaH rather than hydrolysis by residual water
(Figure 11).

Figure 12. Gaseous products (EGA Micro-GC) from the thermal


decomposition of NaH/DMF.

3. Thermal Stability Evaluation of NaH/DMAc. Similar


to the thermal decomposition of NaH/DMF, ARC analysis of
Figure 11. Thermogram of NaH in mineral oil by EGA Micro-GC. a mixture of 9% NaH, 6% mineral oil, and 85% DMAc
evidenced a decomposition onset temperature of 56.4 °C with
The profiles of the gaseous products resulting from the a peak self-heating rate of 5.78 °C/min and a total heat release
thermal decomposition of the NaH/DMF mixture are shown of −422.6 J/g. Increasing the content of NaH to 16.2%
in Figure 12. Water was detected when the temperature was significantly lowered the decomposition onset temperature to
below 140 °C and was likely the result of residual water in the 30.1 °C and substantially increased the maximum self-heating
EGA Micro-GC system, transfer lines, and sample cup. Two rate to 479.1 °C/min with a total heat output of −528.2 J/g
separate peaks were detected for H2, the first at 140−270 °C (Figure 13 and Table 3). Again, both experiments resulted in
and the second at 270−360 °C. In addition to H2, carbon significantly high cool-down pressures, confirming the
monoxide (CO), methane (CH4), and ethylene or acetylene formation of noncondensable gaseous products from the
were also observed within the same temperature range as the thermal decomposition. These results confirm that the thermal
first H2 peak. A trace amount of carbon dioxide (CO2) was decomposition of the NaH/DMAc combination could also
also found at >300 °C.22 The formation of these gaseous result in runway scenarios.
products suggests that a more complicated, potentially radical EGA Micro-GC analysis of the NaH/DMAc reaction
decomposition mechanism occurs during the thermal decom- mixture showed that the release of H2 occurred at a lower
position of NaH/DMF. temperature (<110 °C) than with the NaH/DMF sample, with
E DOI: 10.1021/acs.oprd.9b00276
Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Article

Figure 13. ARC heat rate and pressure vs temperature profiles of NaH/DMAc mixtures.

Table 3. ARC Analysis of NaH/DMAc Mixtures


sample description 9% NaH + 6% mineral oil + 85% DMAc 16.2% NaH + 10.8% mineral oil + 73.1% DMAc
total sample mass (g) 4.2240 5.3631
Cp of sample (J g−1 °C−1) 1.975 1.943
ARC cell mass (g) 22.0010 14.7578
set end temperature (°C) 350 200
phi 2.11 1.60
onset temperature (°C) 56.4 30.1
peak temperature (°C) 123.0 122.7
end temperature (°C) 157.8 200
max self-heating rate (°C/min) 5.78 479.1
total heat (J/g) −422.6 −528.2

the majority of H2 generation occurring in the temperature explosion during the ARC analysis that was forceful enough to
range of 104−305 °C. Other noncondensable gaseous displace the ARC housing. We expect this report to promote
products, such as CH4, ethane, ethylene or acetylene, and awareness of these hazards in the chemistry community,
CO, were also observed in this temperature range as well as encouraging scientists to employ safer alternative conditions
above 300 °C (Figure 14).22 (e.g., alternative bases, NaH with stable solvents, etc.)23 and
avoid incidents related to these hazards.

■ EXPERIMENTAL SECTION
General. All reagents were commercially available and used
as purchased without further purification.
Procedure for ARC Analysis. An accelerating rate
calorimeter manufactured by Thermal Hazard Technology
was used in this study. A preweighted NaH/mineral oil sample
was loaded into a Hastelloy C ARC cell. The ARC sphere was
connected to the calorimeter lid and purged with N2. The ARC
test was started and put in EXO mode, and then the reaction
solvent was manually injected with a syringe. The ARC
Figure 14. Gaseous products (EGA Micro-GC) from the thermal experiment was performed in heat−wait−search (HWS)
decomposition of NaH/DMAc. mode. A heat step of 5 °C, waiting time of 30 min, and
detection threshold of 0.02 °C/min were utilized for each

■ CONCLUSIONS
The inherent thermal instability hazards of NaH/DMSO,
sample.
Procedure for DSC Analysis. A Q2500 differential
scanning calorimeter from TA Instruments was used for the
NaH/DMF, and NaH/DMAc mixtures have been well- constant-heating-rate tests (at 10 °C/min) in this study. A
documented in the literature over the past few decades, yet 3.2515 mg sample (a mixture of 10.3% NaH, 6.9% mineral oil,
the dangers remain underappreciated and poorly communi- and 82.8% DMSO) was sealed within a gold-plated high-
cated, as evidenced by their continued appearance in hundreds pressure crucible with N2 in the headspace. The sealed crucible
of publications. The data outlined in this contribution confirm had a total internal volume of ∼20 μL and was able to
that these reactive mixtures undergo exothermic decomposi- withstand pressures of up to 3000 psi at 400 °C, preventing the
tion at relatively low temperatures, occurring concurrently with release of any reaction materials, products, or byproducts. The
the generation of noncondensable gases. The dangers DSC analysis of pure DMSO was performed following the
associated with NaH/DMSO mixtures were evidenced by an same method in a sealed glass capillary tube under N2.24
F DOI: 10.1021/acs.oprd.9b00276
Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Article

Procedure for EGA Micro-GC Analysis. A Varian model Channel 2: 10 m MS5A module with backflush, SN 61162, PN
CP-490 Micro-GC system was used for EGA of the NaH/ 494001360. Channel 3: 10 m PPQ module with backflush, SN
solvent mixtures using a Frontier Py2020iD pyrolysis unit 61222, PN 494001430. Channel 4: 6 m 19CB module, SN
(Figure 15). Approximately 2 mg of NaH (60% dispersion in 61607, PN 492004810. Dell Latitude Laptop PC with Agilent
EZChrom Elite software, version 3.2.2 SP2.
Method Settings. Sample time: 40 s. Sample line temper-
ature: 100 °C. Stabilizing time: 5 s. Continuous flow: disabled.
Number of flush cycles: 1. Peak simulation: disabled. The
channel settings are shown in Table 4.
Procedure for Headspace GC/MS Analysis. The sample
was preheated at 40 °C for 1 h in a headspace vial and analyzed
along with an empty headspace vial, which served as the blank.
An Agilent 5975C inert XL GC/MSD system with triple-axis
detector was used. The instrument was equipped with both
direct liquid, SPME, and headspace capabilities. A Supelco Q-
PLOT capillary column with dimensions of 30 m × 0.32 mm
(cat. no. 24242) was used for the analysis. The agitation
Figure 15. Flow diagram of the EGA Micro-GC system. temperature of the sample compartment was set at 40 °C. The
NIST chemical library was used for the identification of
mineral oil) was added with 15 μL of solvent (DMF or gaseous products.
DMAc). The sample was analyzed by EGA Micro-GC using
helium as the purge gas through the pyrolysis oven. The ramp
rate of the pyrolysis oven was 5 °C/min, and the samples were
■ AUTHOR INFORMATION
Corresponding Author
heated from 40 to 400 °C. The Micro-GC system was used to *E-mail: qiang.yang@corteva.com.
analyze the gases evolved after the sample was inserted into the ORCID
pyrolysis oven and the programmed heating cycle was started.
Qiang Yang: 0000-0003-3762-5015
An aliquot of the pyrolysis effluent was injected into the Micro
Siyu Tu: 0000-0002-6126-5729
GC system every 150 s. The helium flow rate was
approximately 13 mL/min. Author Contributions
A Molsieve 5A (MS5A) module with argon as the carrier gas The manuscript was written through contributions of all
was used to separate H2. An MS5A module with helium as the authors. All of the authors approved the final version of the
carrier gas was used to separate CO and CH4. A PoraPLOT Q manuscript.
(PPQ) module with helium as the carrier gas was used to Notes
separate CO2, CH4, water, ethane, and ethylene/acetylene. A The authors declare no competing financial interest.
19CB module with helium as the carrier gas was used to
separate DMF and DMAc. Quantitation was performed using a
microthermal conductivity detector (TCD). H2, N2, CO, CO2,
■ ACKNOWLEDGMENTS
The authors thank Dr. Gregory T. Whiteker for his
CH4, ethane, propane, and butane responses were calibrated constructive suggestions during the preparation of this
against certified gas standard mixtures (Matheson Tri-Gas). manuscript.
H2O responses were identified and estimated on the basis of
literature references.25
Varian CP-490 Quad-Channel Micro-GC Specifications.
■ REFERENCES
(1) Hansley, V. L.; Carlisle, P. J. The Production of Sodium Hydride
Serial no. (SN): GC0912B639. Channel 1: 20 m MS5A and Some of its Reactions. Chem. Eng. News 1945, 23 (15), 1332−
module with backflush, SN 61685, part no. (PN) 494001370. 1380.

Table 4. Micro-GC Channel Settings

channel 1 channel 2 channel 3 channel 4


module 20 m MS5A 10 m MS5A 10 m PPQ 6 m 19CB
carrier gas argon helium helium helium
column temp. (°C) 146 90 73 80
injection temp. (°C) 100 100 100 100
injection time (ms) 50 50 50 50
backflush time (s) 10 10 20 NA
detector on on on on
TCD temp. limit check on on on on
sensitivity auto auto auto auto
pressure mode static static static static
initial pressure (psi) 30 30 16 18
sampling frequency (Hz) 100 100 100 100
run time (s) 150 150 150 150
acquisition delay (s) 0 0 0 0

G DOI: 10.1021/acs.oprd.9b00276
Org. Process Res. Dev. XXXX, XXX, XXX−XXX
Organic Process Research & Development Article

(2) Encyclopedia of Reagents for Organic Synthesis; John Wiley & Hydride on Scale: The Process Development of a Chloropyrimidine
Sons, 1995; Vol. 7, pp 4568−4571. Displacement. Org. Process Res. Dev. 2011, 15, 1442. (c) Yang, Q.; Li,
(3) (a) Bretherick, L. Dimethyl sulfoxide: Sodium hydride. In X.; Lorsbach, B. A.; Muhuhi, J. M.; Roth, G. A.; Gray, K.; Podhorez,
Bretherick’s Handbook of Reactive Chemical Hazards, 3rd ed.; D. E. Development of a Scalable Process for the Insecticidal
Butterworths: London, 1985; p 294. (b) Bretherick, L. Sodium Candidate Tyclopyrazoflor. Part 2. Fit-for-Purpose Optimization of
hydride: Dimethylformamide. In Bretherick’s Handbook of Reactive the Route to Tyclopyrazoflor Featuring [3 + 2] Cyclization of 3-
Chemical Hazards, 3rd ed.; Butterworths: London, 1985; p 1133. Hydrazinopyridine·2HCl and Methyl Acrylate. Org. Process Res. Dev.
(4) (a) Corey, E. J.; Chaykovsky, M. Methylsulfinylcarbanion. J. Am. 2019, DOI: 10.1021/acs.oprd.9b00128.
Chem. Soc. 1962, 84, 866−867. (b) Corey, E. J.; Chaykovsky, M. (24) (a) Tou, J. C.; Whiting, L. F. A cradle-glass ampoule sample
Methylsulfinyl Carbanion (CH3-SO-CH2−). Formation and Applica- container for differential scanning calorimetric analysis. Thermochim.
tions to Organic Synthesis. J. Am. Chem. Soc. 1965, 87, 1345−1353. Acta 1980, 42, 21. (b) Yang, Q.; Canturk, B.; Gray, K.; McCusker, E.;
(5) (a) Price, G. G.; Whiting, M. C. The sodium derivative of Sheng, M.; Li, F. Evaluation of Potential Safety Hazards Associated
dimethyl sulfoxide, Dimsylsodium. Chem. Ind. (London) 1963, 775− with the Suzuki−Miyaura Cross-Coupling of Aryl Bromides with
776. (b) Sjoberg, K. Stable solutions of methylsulfinyl carbanion. Vinylboron Species. Org. Process Res. Dev. 2018, 22, 351−359.
Tetrahedron Lett. 1966, 7, 6383−6384. (c) Lyness, W. I.; O’Connor, (c) Yang, Q.; Cabrera, P. J.; Li, X.; Sheng, M.; Wang, N. X. Safety
D. E.; Berry, J. S. Sulfinyl carbanions for preparing surface-active Evaluation of the Copper-Mediated Cross-Coupling of 2-Bromopyr-
detergents. U.S. Patent 3,288,860, Nov 29, 1966. idines with Ethyl Bromodifluoroacetate. Org. Process Res. Dev. 2018,
(6) French, F. A. Sodium Hydride−DMSO Mixture Explodes. Chem. 22, 1441−1447. (d) Yang, Q.; Sane, N.; Klosowski, D.; Lee, M.;
Eng. News 1966, 44 (15), 48. Rosenthal, T.; Wang, N. X.; Wiensch, E. Mizoroki−Heck Cross-
(7) Russell, G. A.; Weiner, S. A. Methylation of Aromatic Coupling of Bromobenzenes with Styrenes: Another Example of Pd-
Hydrocarbons by Dimethyl Sulfoxide in the Presence of Base. J. Catalyzed Cross-Coupling with Potential Safety Hazards. Org. Process
Org. Chem. 1966, 31, 248. Res. Dev. 2019, DOI: 10.1021/acs.oprd.9b00126. (e) Sheng, M.;
(8) Olson, G. L. Lab explosions. Chem. Eng. News 1966, 44 (24), 7. Valco, D.; Tucker, C.; Cayo, E. Practical Use of Differential Scanning
(9) Powers, J. C.; Seidner, R.; Parsons, T. G. The cleavage of formyl Calorimetry for Thermal Stability Hazard Evaluation. Org. Process Res.
groups by sodium hydride. Tetrahedron Lett. 1965, 6, 1713−1716. Dev. 2019, DOI: 10.1021/acs.oprd.9b00266.
(10) Brimacombe, J. S.; Jones, B. D.; Stacey, M.; Willard, J. J. (25) Dietz, W. A. Response factors for gas chromatographic analyses.
Alkylation of carbohydrates using sodium hydride. Carbohydr. Res. J. Chromatogr. Sci. 1967, 5, 68−71.
1966, 2, 167.
(11) Nasipuri, D.; Bhattacharyya, A.; Hazra, B. G. Novel
Aromatisation Reaction of Cyclohexenone Derivatives with Sodium
Hydride. J. Chem. Soc. D 1971, 13, 660.
(12) Buckley, J.; Webb, R. L.; Laird, T.; Ward, R. J. Report on
thermal reaction. Chem. Eng. News 1982, 60 (28), 5.
(13) DeWall, G. Sodium hydride and DMF. Chem. Eng. News 1982,
60 (37), 5.
(14) (a) Laird, T. Org. Process Res. Dev. 2002, 6, 876. (b) Laird, T.
Org. Process Res. Dev. 2005, 9, 951.
(15) Dahl, A. C.; Mealy, M. J.; Nielsen, M. A.; Lyngsø, L. O.; Suteu,
C. Route Scouting and Process Development of Lu AA26778. Org.
Process Res. Dev. 2008, 12, 429.
(16) Ishii, Y.; Fujimoto, R.; Mikami, M.; Murakami, S.; Miki, Y.;
Furukawa, Y. Practical syntheses of Chiral α-Amino Acids and Chiral
Half-Esters by Kinetic Resolution of Urethane-Protected α-Amino
Acid N-Carboxyanhydrides and Desymmetrization of Cyclic meso-
Anhydrides with New Modified Cinchona Alkaloid Catalysts. Org.
Process Res. Dev. 2007, 11, 609.
(17) Stoessel, F. Thermal Safety of Chemical Processes: Risk Assessment
and Process Design; Wiley-VCH: Weinheim, Germany, 2008.
(18) This combination is a result of using 60% NaH in mineral oil
for the evaluations.
(19) Yoshida, T.; Yoshizawa, F.; Itoh, M.; Matsunaga, T.; Watanabe,
M.; Tamura, M. Prediction of Fire and Explosion Hazards of Reactive
Chemicals (I). Estimation of Explosive Properties of Self-Reactive
Chemicals from SC-DSC Data. Kogyo Kayaku 1987, 48, 311−316.
(20) Brandes, B. T.; Smith, D. K. Calorimetric study of the
exothermic decomposition of dimethyl sulfoxide. Process Saf. Prog.
2016, 35, 374−391.
(21) This method was not designed for the detection of hydrogen
gas. It would have coeluted with the carrier gas helium if it was
present as part of the decomposition products.
(22) The presence of dimethylamine (Me2NH) in the decom-
position products has been reported previously.12 Detection of
Me2NH was not attempted during this analysis.
(23) (a) Prashad, M.; Har, D.; Hu, B.; Kim, H.-Y.; Girgis, M. J.;
Chaudhary, A.; Repič, O.; Blacklock, T. J.; Marterer, W. Process
Development of a Large-Scale Synthesis of TKA731: A Tachykinin
Receptor Antagonist. Org. Process Res. Dev. 2004, 8, 330−340.
(b) McCabe Dunn, J. M.; Duran-Capece, A.; Meehan, B.; Ulis, J.;
Iwama, T.; Gloor, G.; Wong, G.; Bekos, E. The Safe Use of Sodium

H DOI: 10.1021/acs.oprd.9b00276
Org. Process Res. Dev. XXXX, XXX, XXX−XXX

You might also like