Tianye
Zheng†
ab,
Haihong
Bao†
*ab,
Feifan
Chen†
ab,
Jingwen
Wu
ab,
Pengcheng
Zhao
ab,
Hoi Lut
Ho
ab,
Shoufei
Gao
c,
Yingying
Wang
c,
Jiaqiang
Huang
d,
Leiting
Zhang
e,
Steven T.
Boles
f and
Wei
Jin
*ab
aDepartment of Electrical and Electronic Engineering, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong. E-mail: darren.ty.zheng@connect.polyu.hk; haihong.bao@polyu.edu.hk; wei.jin@polyu.edu.hk
bPhotonics Research Institute, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong
cInstitute of Photonics Technology, Jinan University, Guangzhou, China
dSustainable Energy and Environment Thrust, The Hong Kong University of Science and Technology, Guangzhou, China
eDepartment of Chemistry–Ångström Laboratory, Uppsala University, Uppsala, Sweden
fDepartment of Energy and Process Engineering, Faculty of Engineering, Norwegian University of Science and Technology, Trondheim, Norway
First published on 8th August 2025
Gaseous molecules are inherent byproducts of (electro-)chemical reactions in lithium-ion battery cells during both formation cycles and long-term operation. While monitoring gas evolution can help understand battery chemistry and predict battery performance, the complex nature of gas dynamics makes conventional mass spectrometry approaches insufficient for real-time detection. Here, we present a radically different methodology for operando analysis of gas evolution in lithium-ion batteries using optical fiber photothermal spectroscopy. By placing an optical hollow-core fiber inside the battery cell, evolved gases can rapidly diffuse into the hollow core of the fiber, enabling photothermal spectroscopy which precisely and selectively quantifies their concentrations without altering the internal operation of the cell. This approach facilitates identification of individual gaseous species, thereby allowing for further clarification (electro-)chemical reaction pathways. Collectively, we show that the evolution paths of C2H4 and CO2 are closely associated with the formation of the solid electrolyte interphase, the selection of electrolyte salts, and the inclusion of specific additives. Significantly, we confirm for the first time the spontaneous formation of CO2, which occurs exclusively in the presence of LiPF6 salt. Beyond the scope of batteries, the methodology presented here offers substantial potential for broader applications, particularly in characterizing electrocatalytic processes, providing unmatched precision, accuracy, and scalability compared to existing analytical techniques.
Broader contextWith the wide spread of electric vehicles and stationary energy storage facilities, ensuring the sustainable production and safe operation of rechargeable batteries is of vital importance. A reliable diagnostic tool capable of real-time monitoring of battery cells would be highly beneficial for elucidating cell chemistry and optimizing cell performance. Here, we present a miniature gas sensor based on optical fiber photothermal spectroscopy for operando monitoring of gas evolution inside battery cells. The new technique enables highly selective gas detection and reveals comprehensive gassing dynamics, offering valuable insights into internal chemical processes. Importantly, the sensing platform presented in this work can be readily extended to monitor additional gas species or adapted for use in other electrochemical energy systems. |
Meeting these requirements is challenging for existing technologies. Currently, in academia, operando studies of battery gassing dynamics predominantly rely on the differential/online electrochemical mass spectrometry (DEMS/OEMS), which involves flowing a carrier gas, such as argon or helium, through a battery cell to collect gases and transport them to a mass spectrometer for measurement. While DEMS/OEMS can be sensitive and versatile in quantifying various gaseous species, the equipment is generally cumbersome and may cause errors during the measurements. The system requires a specially designed cell connected to a gas flow channel, which modifies the internal environment of the cell. The introduction of a carrier gas can further alter the partial pressures of gases within the cell. For example, the solubility of CO2 in organic carbonate-based electrolyte solutions may be affected by the partial pressures, thus hindering accurate measurements of CO2.6 Gas intermediates, such as PF5 and POF3,7 may be stripped out from the cell before participating in further parasitic reactions. Additionally, a mass spectrometer identifies gas molecules by measurement of the mass-to-charge ratio i.e., m/z, which suffers from insufficient specificity for gas detection with overlapped m/z signals (e.g., C2H4, CO, and N2).8–10
Optical fiber sensors have been demonstrated for operando measurements of temperature, strain, refractive index (RI), and gas pressure inside LIB cells.5,11–13 These sensors offer advantages of compact structure, intrinsic safety, corrosion resistance, and immunity to electromagnetic interference without impacting the working conditions of LIBs. However, they do not capture key chemical indicators like gas species and concentrations, which are critical for comprehensive analyses of internal (electro-)chemical reactions. As motivated by these recent advancements, we propose employing hollow-core fiber (HCF) photothermal spectroscopy (PTS) sensors originally developed for atmospheric gas analysis for application in battery gas sensing. Relying on the “fingerprint” molecular absorption line of gas molecules, the HCF–PTS sensors have exhibited excellent sensitivity and specificity.14,15 Gas detection down to parts-per-million (ppm) or parts-per-billion (ppb) level can be achieved with a centimeter-long HCF, which could be conveniently embedded into a battery cell for operando gas measurements, therefore providing direct insights into internal (electro-)chemical reactions.2,3,11–13
Here, we report the use of an HCF–PTS sensor for real-time monitoring of the internal dynamics of the common cycling-related gaseous species: C2H4 and CO2, to largely unlock the (electro-)chemical processes in early stages of LIB cell formation. For the first time, we place the micrometer–diameter HCF inside a battery cell, which requires only a minute amount of gas sample for sensing, ensuring fast response and unperturbed internal conditions of the battery cell during the measurement process. The HCF–PTS not only successfully captures the operando evolution of C2H4 with a high selectivity – an electrochemical process that is extensively documented in previous studies, but also enables the detection of a continues CO2 formation – a purely chemical reaction step that has remained elusive with existing gas sensing technologies. Furthermore, the data collected using our HCF–PTS system is compared with those obtained from DEMS/OEMS systems in previous studies, offering a more comprehensive perspective on the internal chemistries and dynamic processes in LIBs, while paving a path for enhancing battery lifetime and sustainability.
A pump laser beam (indicated as an orange arrow) with its wavelength tuned to a specific absorption line of a target gas is delivered into the sensing HCF through a SMF. The gas absorption of the modulated pump inside the HCF generates heating, which modulates the temperature and refractive index (RI) of the gas sample, and hence the phase of a probe beam (indicated as a blue arrow) propagating through the HCF. The photothermal (PT) phase modulation (Δ∅) of the probe beam is given by eqn (1):14,15
Δ∅ ∝ Ct·α(λpump)·L·Ppump | (1) |
Among various fiber-based interferometric configurations developed for detecting the PT phase modulation,14–16 we use the most compact one with the configuration shown in Fig. 1a. The reflections at the SMF-HCF joints forms a low-finesse Fabry–Pérot interferometer (FPI), which detects the phase difference between the reflected probe beams (Fig. 1b). Simultaneous detection of multiple gas species is achieved by using multiple pump lasers, each tuned to the selected absorption lines of the target gas species and modulated at distinct frequencies to generate respective phase modulations. This HCF–PTS gas sensor offers high sensitivity, selectivity, and a large range of measurements. The miniature and electrical-insulating design makes it highly suitable for integration into battery cells.
Schematically illustrated in Fig. 3, the gas detection system of HCF–PTS employs two distributed feedback (DFB) pump lasers with wavelengths tuned to the absorption lines of C2H4 near 1654.5 nm and CO2 near 2004.0 nm. Two optical amplifiers, i.e., a Raman fiber amplifier (RAF) and a thulium-doped fiber amplifier (TDFA), are employed to boost the pump power, thereby enhancing the PT-induced phase modulation and improving the measurement sensitivity. A probe beam from a fiber laser at 1550 nm is combined with the two pump beams using a wavelength-division multiplexer (WDM) and launched into the HCF sensor.
To achieve a higher signal-to-noise ratio (SNR), wavelength modulation and harmonic detection are employed. The pump laser wavelength is modulated sinusoidally at a high frequency and swept slowly across the gas absorption line. For optimal phase detection, the FPI is stabilized at its quadrature point with a servo controller using the DC component of the reflected probe beam detected by photodetector 1 (PD1).21 The reflected probe beam is also detected by PD2 and its second harmonic (2f) signal, which is proportional to gas concentration, is demodulated by a lock-in amplifier (LIA). Frequency-division multiplexing (FDM) is employed to enable the simultaneous detection of multiple gases.22 The two pump lasers are modulated at distinct frequencies, i.e., 10 kHz for C2H4 and 8 kHz for CO2, allowing the 2f signals to be demodulated simultaneously at their corresponding second harmonic frequencies by the LIA. The time constant used in LIA is 400 ms with 18 dB/Oct slope and the scanning frequency of both pump wavelengths is 10 mHz.
The HCF sensor is packaged using a stainless steel (SS) tube and attached to a homemade cell via a Swagelok connector as shown in Fig. 3. The inner volume of the 5-cm-long HCF is ∼0.12 μL, while the total volume of the battery cell is estimated to be ∼1.4 mL, resulting in a minimal influence on the gas environment within the cell. To monitor the overall change of the gas pressure within the LIB cell during charge–discharge, a commercial piezoelectric pressure sensor is also incorporated. After installing the gas and pressure sensors, cyclic voltammetry (CV) measurements are conducted, and results indicate that the battery performance remains largely unaffected. Detailed data can be found in Note 2, SI. All measurements were conducted with the homemade cell placed in an environmental chamber maintained at a constant temperature of 35 °C.
![]() | (2) |
Additionally, it is necessary to characterize the performance of our HCF–PTS gas detection system in terms of response time and limit of detection (LOD) using standard gas samples under conditions of 35 °C and 1 atm. The response time of the HCF sensor, defined as the time required for the signal to rise from 0 to 90% of its maximum value (t90),14 is measured to be 14.5 seconds. The LOD is determined to be 2.5 ppm for C2H4 and 1 ppm for CO2. Detailed determinations of response time and LOD are provided in Note 4, SI.
The rates of change in the gas concentrations and pressure can then be determined by calculating their time derivatives, i.e., dC/dt and dP/dt, aiming to shed light on the gassing events occurring at specific time points and/or electrode potentials. Fig. 4e presents the GCD profiles of the LFP-graphite full cell, along with the rates of change for overall pressure (dP/dt), as well as concentrations of C2H4 (dCC2H4/dt) and CO2 (dCCO2/dt). From these results, the HCF–PTS system successfully captures well-documented gassing events, where the dCC2H4/dt profile shows a gassing onset at a cell voltage of ∼2.8 V (equivalent to an anode potential of ∼0.6 V vs. Li/Li+), about 20 minutes after the charging starts and reaches a peak formation rate at dCC2H4/dt ≈ 6000 ppm h−1. The C2H4 evolution stops at ∼3.3 V (equivalent to an anode potential of ∼0.1 V vs. Li/Li+) and does not appear in the subsequent cycle, indicating that the selected initial charging rate should be sufficiently slow to form a passivating SEI layer on the graphite surface. During the initial charging of an LIB cell, C2H4 evolution is inevitable when EC is used as one of the electrolyte solvents, as it has been reported to originate exclusively from EC reduction with a lithium ethylene decarbonate (LEDC) intermediate electro-reduction product.19,24,25 After the initial charging, the C2H4 concentration seems to decrease very slowly, from approximately 0.78% to 0.71% by the end of the second cycle (Fig. 4c). A similar trend is also observed in other studies using the OEMS technique.26,27 Ellis et al. concluded that the C2H4 formed during the initial charging is gradually consumed at the cathode evidenced by an increase in cell impedance.28 The small time-lag in the onset of the dCC2H4/dt profile, compared to that in the dP/dt profile, can be attributed to the possible evolution of H2, which likely originates from the reduction of protic electrolyte oxidation species (R–H+) from the cathode.26 Overall, the observed C2H4 dynamics correlate well with the dP/dt profile and are consistent with the previous studies,24,26–28 affirming the validity of the measurements from the HCF–PTS system.
Beyond confirming previously reported C2H4 dynamics, our system also uncovers new and intriguing dynamics of CO2 evolution: Before charging the LIB cell, the CO2 concentration is non-zero and keeps increasing at a rate of dCCO2/dt ≈ 200–300 ppm h−1. The non-zero level of CO2 may partly originate from the background CO2 in the glovebox, which was verified by a commercial electrochemical CO2 sensor. The increase in CO2 concentration before cycling suggests spontaneous CO2-forming reactions without electrochemical driving forces. This spontaneous CO2 formation is further confirmed by monitoring the gas evolution in the cell that only consists of the electrolyte and separator (Note 6, SI). Additionally, the airtightness of the homemade cell was double-checked to prove the claim (Note 7, SI). Once the battery charging begins, dCCO2/dt decreases significantly. When the cell voltage rises from 2.8 V to 3.3 V, the CO2 concentration decreases with a maximum rate of dCCO2/dt ≈ −900 ppm h−1. By integrating dCCO2/dt over time, the total decrease in CO2 concentration is determined to be 1832 ppm as indicated by the shaded area below zero in the dCCO2/dt profile shown in Fig. 4e. As the cell voltage rises beyond 3.3 V, dCCO2/dt gradually restores to ∼200–300 ppm h−1 after the C2H4 evolution concludes. Notably, the dCCO2/dt profile remains stable and does not vary significantly during subsequent cycling, with the CO2 evolution being independent of the cell voltage. Such spontaneous CO2 formation in the absence of electrochemistry was not clearly acknowledged in previous studies using the DEMS/OEMS systems.6,26,27,29,30 At the end of the second charge, a decrease in dCCO2/dt is observed as marked by a dashed arrow in the 3rd column of Fig. 4e when the cell voltage is above ∼3.5 V. Unlike the concentration decrease observed during the first charge, this decrease does not coincide with C2H4 generation, which suggests a different mechanism behind the occurrence of CO2 consumption, (re-)dissolution, or other gas-forming reactions that dilute the CO2 concentration. Collectively, the newly revealed gassing dynamics from the classic Li-ion battery system merits further investigation and will be discussed in the following sections.
Fig. 5a presents the initial CV curve obtained from the Li-graphite cell alongside the evolution profiles of both C2H4 and CO2 gases, as well as the overall gas pressure. Notably, the profile of dCC2H4/dt aligns well with the dP/dt, indicating a strong correlation between the C2H4 dynamics and the change in overall gas pressure. A distinct cathodic bump appears with an onset potential of ∼0.9 V, peaking just below ∼0.8 V vs. Li/Li+ as shown in the inset of the current–potential plot. This bump synchronizes with the increase of overall gas pressure and the initiation of C2H4 evolution results from the reduction of EC on the graphite electrode near ∼0.8 V vs. Li/Li+.26,27 In contrast, the dCCO2/dt profile shows that CO2 concentration increases at a rate of approximately 200 ppm h−1 immediately after cell assembly. Interestingly, this spontaneous CO2 formation does not significantly impact the overall gas pressure (≤0.1 kPa h−1), suggesting that CO2 solubility-vapor pressure dynamics are likely involved. As shown in Fig. 5b, the electrochemical impedance spectroscopy (EIS) measurements reveal no significant difference in cell impedance after the initial and the second cathodic scan, indicating the formation of a stable passivating SEI. Following the SEI formation, a pronounced increase in dCCO2/dt to 400 ppm h−1 is observed, during which no C2H4 generation or notable pressure variation occurs, suggesting that C2H4 and CO2 evolution may be decoupled, with CO2 also involved in other electrolyte decomposition reactions.25
Fig. 5c presents an alternative plot of gas concentration vs. capacity for the initial cathodic scan, where the C2H4 concentration remains negligible until shooting up at nearly 0.1 mAh. Meanwhile, the CO2 profile appears independent of capacity, suggesting that its formation is through spontaneous reaction pathways. By integrating the cyclic voltammograms in the inset of Fig. 5a (blue shaded areas), we estimate the total electrons that are associated with the SEI formation until 0.3 V vs. Li/Li+ to be approximately 6.84 μmol (Note 9, SI). At this electrode potential, the intercalation of Li into graphite has not yet started. These electrons are irreversibly consumed in the formation of the SEI layer and the generation of gases, leading to an ICE of ∼86%. By applying the ideal gas law to the recorded pressure data, the change in total gas molecules within the cell is estimated to be ∼0.64 μmol. Meanwhile, our HCF–PTS gas sensor quantifies the C2H4 molecules to be ∼0.68 μmol through multiplying the concentration with the cell volume (ca. 1.4 mL). These quantitative results confirm that C2H4 is the primary gas that evolved during SEI formation, playing a significant role in influencing overall cell pressure, aligning well with previous investigations.24,28,29
In contrast, CO2 does not substantially influence the overall gas pressure. The high solubility of CO2 in the electrolyte allows it to dynamically dissolve into or escape from the electrolyte, buffering changes in the overall gas pressure. Additionally, it is suggested that the CO2 formation observed in Fig. 5a may be linked to the presence of LiPF6. The LiPF6 salt dissolved in the carbonate solvents can readily dissociate to form gaseous PF5,7,32,33 which is highly reactive with cyclic carbonate solvents, such as EC, triggering spontaneous CO2 generation. This was evidenced by the observation that CO2 generation is accompanied by a decrease of EC in an electrolyte of LiPF6 in EC:DMC (dimethyl carbonate) at 85 °C.34,35 Similarly, Ravdel et al. employed both nuclear magnetic resonance (NMR) spectroscopy and gas chromatography-mass spectrometry (GC-MS) and observed that the dissociated PF5 gas reacts with EMC, producing gaseous CO2 and POF3.36 Further NMR studies indicate that POF3 may cleave EC, resulting in the formation of CO2 and OPF2OCH2CH2F. This intermediate subsequently reacts with additional EC, further liberating CO2 and forming capped oligoethylene oxides [OPF2(OCH2CH2)nF] by inserting ethylene oxide units into the P–O bond.37 Our results suggest that these CO2 formation reactions are largely independent of cell potential and are likely spontaneous and dependent on CO2 presence itself. Scheme 1 summarizes the potential reaction pathways for spontaneous CO2 formation, drawing on insights from prior studies.
![]() | ||
Scheme 1 Suggested reaction pathways for the spontaneous CO2 formation in LiPF6//carbonate electrolytes. |
The higher dCCO2/dt rate observed after the initial cathodic scan is perhaps caused by the decompositions of certain SEI components, such as Li2CO3, via LiPF6 + Li2CO3 → CO2(g) + 3LIF + POF3(g).19 Lithium methyl carbonate (LMC) may also thermally decompose to form CO2 in the presence of LiPF6 at 55 °C via LiPF6 + 3LMC → 3CO2(g) + 4LiF + OPF2OCH3 + CH3OCH3.4 Parimalam et al. found that adding LiPF6 leads to the decompositions of certain SEI components during cell storage, resulting in the generation of a mixture of CO2, LiF, ethers, phosphates, and (fluoro)phosphates. These results were confirmed by multiple experimental tools, including NMR, GC-MS, and infrared spectroscopy with attenuated total reflectance (IR-ATR).4 Additionally, due to the difficulty in completely eliminating residual moisture from the electrodes and electrolytes, trace hydrofluoric acid (HF) may present in the LiPF6-based electrolyte, which decomposes lithium alkyl carbonate to form CO2via LiOCO2CH3 + HF → LiF+ + CH3OH + CO2.8
According to literature, non-spontaneous (i.e., electrochemically-driven) CO2 formation often occurs at the cathode surface with a high potential vs. Li/Li+ due to oxidation of organic carbonate solvents via RCO3R → ROR + CO2(g).8,38–40 When a cell is overcharged, the oxidation of conductive carbon and electrode binder may also lead to CO2 formation.29 However, neither of these cases seems to be occurring in the Li-graphite half-cell, where the electrode potential is relatively low. It is also reported that a ‘water-contaminated’ cell (∼300 ppm H2O) exhibits a more pronounced CO2 evolution than a ‘water-free’ cell (≤20 ppm H2O) when 1 M LiPF6 in EC:DEC (diethyl carbonate) is used as the electrolyte.6 An extra experiment was conducted to investigate the effect from water contamination, affirming and supporting that our drying process is sufficient. When the cell parts were not dried aggressively, the dCCO2/dt profile alters markedly with a notably higher level of CO2 concentration (details in Note 10, SI).
The insignificant change of the dCC2H4/dt profile upon replacing the electrolyte salt is consistent with the observation that C2H4 evolution results from the reduction of EC at the graphite electrode, which remains unchanged in both electrolytes. This claim is further supported by the EIS spectra in Fig. 6b, showing no noticeable change in cell impedance, thereby indicating similar electrochemical properties of the LiClO4-derived SEI, which however does not contain inorganic species like LiF. The C2H4 dynamic shown in Fig. 6c is consistent with those of the cell using the LiPF6-based electrolyte, though with a slightly shifted onset of increase of ≤0.1 mAh due to the elimination of the possible CO2-producing reactions. Generally, this allows us to conclude that the series of reactions leading to C2H4 generation is PF6− independent, consistent with the EC-LEDC mediation discussed above.
Regarding CO2 formation, while the precise reaction pathways can be quite complex, we provide strong evidence that its spontaneous formation originates from the presence of PF6−,36,37 aligning well with the reaction pathways proposed in Scheme 1. In the LiClO4-based electrolyte where PF6− is absent, no spontaneous CO2 generation is detected either before or after the formation of SEI. Notably, CO2 concentration decreases at the start of the CV scan. The total decrease in CO2 concentration is determined to be 819 ppm by integrating dCCO2/dt over time as electrode potential decreases from 1.8 V to 1.0 V, suggesting that the background CO2 in the glovebox may be dissolved into the carbonate electrolyte after cell assembly. Generally, this finding after replacing the electrolyte salt is consistent with a recent study by Lundström et al., who observed no clear sign of CO2 evolution with LiClO4 as the salt and EC as the solvent.30
The negligible level of dCCO2/dt during the second cathodic scan again underscores the role of LiPF6 in facilitating the spontaneous CO2 formation via both the electrolyte breakdown and SEI decomposition discussed above. While this LiClO4-based electrolyte experiences less pressure change and negligible CO2 production in the formation cycle, the presence of fluorine to passivate the aluminum-based cathode current collector mandates the use of LiPF6-based electrolytes in commercial cells.
In our experiment, the overall pressure change is too small to be detected by the commercial piezoelectric pressure sensor after adding VC. This is attributed to the more stable VC-derived SEI, which inhibits reduction reactions on the graphite anode surface and leads to less intensive gassing activity.27 The results demonstrate that this minimal pressure change correlates with a significantly lower rate of dCC2H4/dt, peaking at just above 1000 ppm h−1, which is notably smaller than those shown in Fig. 5a and Fig. 6a. The significantly lower dCC2H4/dt in the presence of 2 vol% VC is attributed to the early reduction of VC at a relatively higher potential vs. Li/Li+ than that of EC. While the exact onset potential for VC breakdown remains debatable, it is generally accepted to be above 1 V vs. Li/Li+.41–45 The VC-derived SEI layer effectively suppresses EC reduction at lower potentials, consistent with the reduced magnitude of the dCC2H4/dt profile shown in Fig. 7a.
The initial cathodic CV scan shows that both the onset and the position of the first reduction peak slightly shift towards a positive potential with a lower catalytic current (i.e., the reduction peak broadens), as compared to the cell using the VC-free electrolyte (Fig. 5a). The EIS spectra in Fig. 7b also notably altered, exhibiting a larger cell impedance with a higher surface potential at ∼130 mV vs. Li/Li+. Pritzl et al. conducted a quantitative study showing that a small addition of VC could lead to a higher impedance of the graphite anode, but (perhaps counterintuitively,) coupled with better cycling performance. Further addition of VC, however, can result in even higher impedances of both electrodes, leading to poorer cyclability.42 Therefore, an optimal dosage of VC is required to grow a reliable SEI layer that suppresses gas evolution, but without excessively increasing cell impedance.41 In this regard, our gas sensing system can be an ideal tool to determine the optimal dosage of VC (or other electrolyte additives) through correlations between gas concentrations and cell impedances.
Generally, formation and consumption reactions of CO2 can occur simultaneously at voltages below 1.5 V in LIB cells with common LiPF6//carbonate electrolytes,6 and the addition of VC further complicates the interpretation of the dCCO2/dt profile. In one study, VC reduction is reported to occur at a potential as high as ca. 1.9 V vs. Li/Li+ and the generation of CO2 involves VC ring-opening reactions driven by a nucleophilic attack that may be a spontaneous process.30 Interestingly, as shown in Fig. 7c, the C2H4 signal also rises at ∼0.1 mAh, like the observation in Fig. 5c. This suggests that the additional generation of CO2 caused by the presence of 2 vol% of VC does not seem to consume extra electrons before the onset of C2H4, supporting the hypothesis that it is driven by a chemical process rather than an electrochemical one. Despite this, a higher level of irreversible capacity, reflected in a poorer ICE of ca. 79.6%, is observed. This poorer ICE indicates that more electric charges are consumed in forming the VC-derived SEI layer rather than in generating CO2 gas. Notably, VC is known to dominate SEI formation as long as it presents,30 explaining why even a small addition of 2 vol% significantly impacts electrochemical signals, alters gassing behaviors, and leads to higher electrode impedances. These findings highlight the duality of VC, influencing both SEI formation and gas evolution dynamics.
As presented in Fig. 8, the schematic illustrates how we enabled the attachment of our HCF–PTS sensor to the LFP-graphite pouch cell using a three-way valve (details can be found in Methods). Firstly, the soft pouch cell package swells evidently, leading to an unstable and dynamic internal pressure. The overall gas volume accumulated after the initial cycle was estimated to be ∼3 ml using a syringe (Note 11, SI), which is at the same magnitude as that of the same type of Li[Ni1/3Mn1/3Co1/3]O2-graphite pouch cell determined by Archimedes’ method.40 The C2H4 profile exhibits the same onset point at a cell voltage of ∼2.9 V, agreeing with that of the homemade cell. However, the C2H4 concentration increases gradually until reaching ∼8.2% and ∼9.3% after initial charge and discharge, respectively. Importantly, given that the background CO2 in the glovebox is only ∼1500–2000 ppm, the spontaneous CO2 formation is reflected by nearly 4% of CO2 prior to charging the cell. As indicated by the orange arrow, a decline in CO2 concentration coincides with the onset of C2H4 evolution. This decrease is suggested to be caused by the observed C2H4 evolution and the possible H2 evolution reported elsewhere.28 Afterward, the CO2 concentration remains approximately at a constant level, as the C2H4 concentration climbs. A more pronounced increase in CO2 concentration is observed during the discharge, accompanied by the stabilized C2H4 profile. Lastly, it should be noted that the gassing dynamics of a pouch cell are not directly comparable to that of our homemade cell due to the continuous change of overall gas pressure caused by cell swelling.
This demonstration clearly supports the potential of our sensing technology to be applied in commercial battery cells. For instance, our gas sensing technique (which detects gas concentration) and the Archimedes’ method developed by Aiken et al. in 2014 (which measures gas volume)46 may naturally form a great combination that leads to comprehensive analyses of gassing dynamics specifically for pouch cells. Applications for sensing other cell formats can be envisioned pending engineering efforts. Additionally, we have demonstrated a tailored water splitting experiment for operando monitoring of O2 and CO2 (detailed discussion in Note 12, SI), showcasing the substantial potential of our gas sensing technique for broader applications beyond the battery realm, particularly for aqueous-based electrochemical energy systems, such as aqueous batteries and electrolysis cells.
Moreover, measurements of gas evolution using DEMS/OEMS systems may disrupt the typical operating environment of an LIB cell due to their sampling-and-then-measuring approach by introducing a carrier gas flow, leading to possible data misinterpretations. In contrast, the miniature HCF sensor enables operando measurement and operates without the need for carrier gas, preserving the natural working conditions of LIB cells. Our approach enables the operando monitoring of gassing dynamics that more precisely reflects the (electro-)chemical reactions occurring within the LIB cells. Remarkably, our HCF–PTS gas sensing system has effectively identified spontaneous CO2 formation immediately after the LIB cell assembly. Furthermore, our findings provide strong evidence that the PF6 anions in electrolytes are the primary driver of this spontaneous CO2 formation. By replacing LiPF6 with LiClO4 to deprive the intermediate gas components of PF5 and POF3, CO2 formation was almost entirely suppressed. This spontaneous CO2 formation becomes more pronounced when the LiPF6-based electrolyte contains 2 vol% VC, emphasizing its significant impact on CO2 evolution in LIB cells. In any case, such detailed observations are challenging to achieve with the DEMS/OEMS systems alone. As to whether CO2 is problematic or not for near term or long-term cell reliability, it is beyond the scope of this work. But the first step in identifying a problem is to acknowledge that it exists, which is now possible with HCF–PTS.
Looking at the future technological roadmap, the HCF–PTS system can be conveniently upgraded for multi-component gas detection beyond C2H4 and CO2. With the wide transmission window of the HCF, spanning from visible to mid-infrared range, multi-component gas detection inside LIBs is achievable. For example, by using additional pump lasers targeting the absorption lines of O2 at 760 nm (Note 12, SI), H2 at 2122 nm, CO at 4600 nm, and employing frequency- or time-division multiplexing techniques for signal separation, the system can detect simultaneously the five most common gas species during battery formation and operation, i.e., C2H4, CO2, H2, CO, and O2, providing a universal gas sensing solution for LIBs. The technique can also be readily extended to characterizing new materials in LIBs48 and sensing sodium-ion batteries with similar cell design and chemistry.49 From an engineering aspect, only the sensor head might degrade while the rest of the interrogation system remains unchanged. The cost of an HCF sensor head is typically in the range of several US dollars, which makes the sensor head financially competitive and replaceable for optimal performance during long-term operation.
The demonstrative experiment of sensing a commercial Li-ion pouch cell proves that our sensing method is readily adaptable for real-world applications, requiring only engineering efforts for sensor integration. In industries, our fiber sensing technology not only offers a promising real-time monitoring solution for rechargeable batteries during operation but also contributes to sustainability by enabling the evaluation of retired batteries, either for recycling or for extending their second life. By integrating the HCF–PTS sensing system with existing operando material characterization tools, such as electrochemical quartz-crystal microbalance (EQCM) and atomic force microscopy (AFM), the growth of SEI as well as the associated chemical reactions can be better understood. Similarly, the evolution of gases may be of critical importance to other applications such as electrocatalysis, artificial photosynthesis, and ammonia oxidation reactions, offering a wealth of possibilities for fiber-based sensing in electrochemical systems beyond batteries. Additionally, the proposed method has the potential to form synergistic combinations with existing gas sensing techniques, such as DEMS/OEMS, setting the stage for rich insights into the internal chemistry of various battery systems and beyond, particularly the field of electrochemical CO2 reduction. While the reduction reactions convert CO2 into various valuable products, whether CO2 (C: +4) is reduced to form CO (C: +2), C2H4 (C: −2), or CH4 (C: −4) remain largely unclear and uncontrollable. By implementing such an operando gas sensing method, ‘electron-to-gas’ pathways can be established for the given CO2 reduction system, providing a scientific basis for further improvements in safety, reliability, and sustainability of future electrochemical energy systems.
Supplementary information is available: Supplementary Note 1–12; Fig. S1–S16. See DOI: https://doi.org/10.1039/d5ee04211a
Footnote |
† Equal contributions. |
This journal is © The Royal Society of Chemistry 2025 |