1. Introduction
There was a rise of about 0.6 °C in measured global mean surface temperature (GMST) between 1970 and 2000 [
1,
2,
3,
4] and a global warming pause for the period 2000–2015 [
2,
3,
5,
6,
7,
8,
9,
10,
11,
12]. The GMST appears to have risen again in the past few years [
13]. In CO
2-based climate models (e.g., general circulation models—GCMs), which were built primarily on the greenhouse effect of CO
2 as the major culprit of global warming, the re-rise in GMST after 2015 is attributed to increasing CO
2 [
13]. In a recent study, however, this author [
14] showed strong evidence that the post-2015 warming is mainly caused by Arctic amplification of surface warming due to sea-ice loss, without which a global reversal in warming would have occurred since ~2005. This conclusion is drawn from comprehensive observed datasets and statistical analyses of troposphere–stratosphere temperature (T) climatology, global lower-stratospheric T (GLST), global surface air T, sea-ice extent (SIE) and snow-cover extent (SCE), combined with analytical, parameter-free calculations of GMSTs by a quantum physical model of climate change caused by halogen-containing greenhouse gases (GHGs) (hereafter halo-GHGs) [
3] and of GLSTs by the cosmic-ray-driven electron-induced reaction (CRE) model of stratospheric O
3 depletion conspired by anthropogenic chlorofluorocarbons (CFCs) and natural cosmic rays (CRs) [
3,
5,
7,
15]. The CRE model has now been developed into a complete quantitative model of global ozone depletion [
16]. These observed and calculated results have reached a surprizing conclusion that global climate change is human-made but dominantly caused by halo-GHGs [
14].
The above conclusion appears quite odd, especially when the 2021 Nobel Prize in Physics was just awarded to some pioneers of GCMs. However, it is not inconsistent with the conclusions reached in a significant body of literature [
3,
5,
6,
7,
8,
17,
18,
19,
20,
21,
22,
23,
24,
25,
26], which do not agree with CO
2 climate models. In fact, there are long debates on the mechanisms and perspectives of climate change, as documented in the publications or lectures by renowned physicists (e.g., Ivar Giaever [
27]—a 1973 Nobel laureate in physics, Freeman Dyson [
28], William Happer [
8], Steven Koonin [
29], and most recently John Clauser—a 2022 Nobel laureate in Physics, according to his keynote lecture at Quantum Korea 2023) and climate scientists (e.g., Richard Lindzen [
22,
30], Reginald Newell [
20], Sherwood Idso [
21,
23], and John Christy [
24,
25]). These previous investigations by others [
8,
17,
18,
19,
20,
21,
22,
23,
24,
25,
26,
27,
28,
29,
30] and this author [
3,
5,
6,
7,
14] have presented a compelling argument that the assumptions made in CO
2 climate models have to undergo close scrutiny. Notably, the former MIT professor Reginald Newell [
20] wrote in 1979: “The fact that water vapor dominates CO
2 in the radiation budget has been known and discussed for many years (see, e.g., Kondratiev and Niilisk [
17]; Moller [
18]; Zdunkowski et al. [
19]) but it seems important to reemphasize when so much attention is being paid to CO
2”. Given with substantial observations found and reviewed [
3,
14], this author feels that such a statement seems not out of date.
This paper has the main purpose of using a series of key observations to assess the CO2-based GCMs and the CFC-warming physical model [
3,
14].
GCMs are a class of ‘
quasi-realistic’ climate models. GCMs not only include unresolved terms represented in equations with
tunable parameters but also have major limitations such as the structural error and uncertainty across models with different representations of unresolved scales and the requirement of tuning the models to match certain aspects of observations [
31]. Despite these weaknesses, GCMs have been a mainstay of climate research for several decades, and the awarding of the 2021 Nobel Prize in Physics to GCM-based work was just praised by the climate community [
32]. As stated in a Perspective published in PNAS in 2022 by Balaji et al. [
31], however, there is current debate on the obsolescence of GCMs, and considerable literature argues that the limitations of GCMs require a major overhaul for their continuing role in climate modeling. The bounds of uncertainties on equilibrium climate sensitivity (ECS) (the equilibrium change in GMST that would result from a doubling of CO
2 concentration) given by GCMs are known to be very large and have not significantly diminished since the late 1970s [
26,
33,
34]. Satellite observations are unable to constrain GCMs adequately, and other indirect means (e.g., paleoclimate data) are often used to place limits on ECS. The most recent IPCC AR6 concluded that many GCMs were providing ECS outside the “very likely” range, and used emulators where ECS is a tunable parameter, to refine the consensus projections and their uncertainty bounds [
31]. These facts imply a diminishing or limited role of GCMs in the past decades (from phase 3 to 6 of the Coupled Model Intercomparison Project, CMIP3 in the IPCC AR4 [
1] to CMIP6 in the AR6 [
13]).
In contrast to GCMs, the author has developed a
conceptual quantum physical model of climate change, which evolved from the findings in studies of stratospheric ozone depletion by the CRE model [
3,
5,
7,
15,
16,
35,
36,
37,
38]. First, it was robustly demonstrated that the variations in both ozone and lower-stratospheric temperature (LST) are dominantly governed by both (halogen-containing) ozone-depleting substances (ODSs) and CR intensity only [
3,
5,
7,
15,
16,
38]. This has in fact revealed no sign of the GCM-predicted greenhouse effect of increasing CO
2 on the stratospheric climate. Second, the author made the first prediction in 2008–2009 that the long-term change in GMST since 1950 was mainly caused by halo-GHGs (mainly CFCs). This prediction was first based on the found correlation between observed GMST and total concentration of atmospheric CFCs [
5] and then on the calculations of changes in GMST due to anthropogenic emission of halo-GHGs [
6] for the period from 1950 to 2009. The subsequent analysis of substantial observed datasets showed strong evidence of the warming ‘hiatus’ associated with declining halo-GHGs regulated by the Montreal Protocol [
3,
7]. Third, a simple quantum physical model of climate change caused by halo-GHGs was finally developed on the basis of the quantum physics of the Earth’s blackbody radiation [
3]. Of special note is that this physical model includes
no tunable parameter and directly gives
analytical calculations of GMST, while exhibiting excellent agreement with observations [
3], particularly after the removal of natural El Niño southern oscillation (ENSO) and volcanic effects [
14]. These results have provided compelling evidence for the dominant warming mechanism of halo-GHGs (mainly CFCS) in the late 20th century [
3,
14]. This CFC-warming physical model has well explained the ‘hiatus’ in global warming and has predicted a long-term
cooling trend starting around the turn of the century, corresponding to the changing trend of atmospheric halo-GHGs.
The ‘hiatus’ in warming since the turn of the century has widely been reported and there exist large discrepancies between observed and GCM-simulated GMST trends in 1998–2015 [
2,
3,
5,
6,
7,
8,
9,
10,
11,
12,
39]. All historical simulations by GCMs did not reproduce the ‘hiatus’ in global warming over the period 2000–2015 [
2,
3,
9,
39,
40]
. Various mechanisms (mainly natural climate variability) have been proposed for the warming ‘hiatus’ [
2,
8,
9,
10,
11,
12,
40], but natural climate variability has its limits. In fact, model simulations gave a duration of an observed 15-year absence of warming as an indication of a significant discrepancy with the prediction of CO
2 climate models and thus predicted in 2009 that “no sort of natural variability could hold off greenhouse warming much longer” and “(rapid) warming will resume in the next few years” [
40].
The change in stratospheric T mirrors the change in surface T. Radiosonde- and satellite-measured data showed that the troposphere had warmed and the stratosphere had cooled from the mid-1970s to the turn of the century [
2,
4]. In CO
2 climate models, the observed cooling of the middle and upper stratosphere was primarily attributed to rising concentrations of well-mixed GHGs (WMGHGs) (mainly non-halogen GHGs) [
41,
42], whereas the changes in LST are attributed to WMGHGs and stratospheric ozone depletion [
2,
4,
43]. CO
2 climate models have also made an iconic prediction that tropospheric warming and stratospheric cooling would continue due to continued WMGHG increases over the coming decades [
2,
13]. In contrast to this prediction and the reported post-2015 rise in surface-measured GMST, however, most satellite measurements have shown that stratospheric Ts steadily decreased to the end of the 20th century, but since then the GLST has stabilized [
14,
43], as widely stated in the 2018 WMO report [
4] and the 2021 IPCC AR6 [
13]. In fact, it is generally agreed that the GLST has dominantly been controlled by ODSs [
3,
4,
5,
7,
14,
15]. Interestingly, a recent analysis of upper-air radiosonde T profiles from 1976 to 2015 by a large international team (Philipona et al. [
43]) has provided remarkable evidence that after decades of cooling, the lower stratosphere at altitudes between 15 and 30 km and over most continents has now been
warming since the turn of the century. They also presented NOAA’s MSU/AMSU satellite data from 1979 to 2015 at the radiosonde locations, showing minor warming in the early 21st century. This reversal in LST from cooling in the late 20th century to current warming has been confirmed in our recent study [
14], using the high-quality EUMETSAT’s ROM SAF satellite datasets of the troposphere-stratosphere T climatology. Furthermore, this author [
14] also discovered that such a reversal has also occurred in the upper stratosphere at altitudes above 30 km over the past decade, especially at high latitudes. This critical observation provides a striking fingerprint that
the radiative (greenhouse) effect of anthropogenic GHGs has been decreasing since the turn of the century.
In this review, the author will first briefly review the fundamental physics of global warming theory in
Section 2. Then, brief reviews of CO
2 climate models and their applications for early calculations of warming caused by halo-GHGs will be given in
Section 3 and
Section 4, respectively.
Section 5 will present a review of
six key observations providing the fingerprints of the contribution of the greenhouse effect of halo-GHGs versus non-halogen GHGs (mainly CO
2) to climate change. A refined quantum physical model of climate change and parameter-free analytical calculations of GMST changes through the CFC-warming mechanism will be reviewed in
Section 6, which is followed by
Section 7 to give remarks and perspectives on global climate change. It will finally address the potential ramifications of this study on the recent proposals of “geoengineering” (stratospheric aerosol injection) to reduce global warming [
44] and of a future generation of GCMs to “match them better to observations, theory, and process models” [
31].
2. Fundamental Physics of Global Warming Theory
The thermal radiation from the Earth or the Sun can approximately be treated as blackbody radiation at temperature T. The radiation intensity spectrum
I(
λ) is obtained by integrating the radiation energy flux per unit wavelength given by Planck’s formula in quantum physics over a certain wavelength interval
Here, the theoretical radiation intensity
I(
λ) spectrum, the satellite-measured outgoing long-wave radiation (OLR) transmittance (Γ) spectrum at the top of atmosphere (TOA) [
45], and the atmospheric transmittance spectrum measured at the ground with the Sun as a blackbody source on Mount Miron at 30° ground-to-space slant path and simulated by Modtran4 [
46] are plotted in
Figure 1 [
3,
6,
7]. The spectral region of 8–13 μm is called the ‘atmospheric window’ as the unpolluted atmosphere in this spectral region is quite transparent, except for absorption by ozone in the 9.6 μm band. Moreover, over 80% of the total radiation energy from Earth’s surface and clouds is emitted into space through this window. As a result, any polluting molecule that strongly absorbs radiation in the atmospheric window is highly effective GHGs. Unfortunately, many halogenated gases such as CFCs and HCFCs have strong absorption bands in this atmospheric window and are therefore not only ODSs but highly effective GHGs. So far, measured halo-GHGs have included CFCs, hydrochlorofluorocarbons (HCFCs), hydrofluorocarbons (HFCs), perfluorocarbons (PFCs), and SF6. CFCs and HCFCs are both ODSs and GHGs, whereas HFCs, PFCs, and SF6 are non-ODSs but GHGs. The major anthropogenic halo-GHGs were CFCs up to the 2000s, while since then the contributions of HCFCs, HFCs, and PFCs to the greenhouse effect have been increasing [
2,
4,
13].
As also shown in
Figure 1, CO
2 contributes to the strong infrared (IR) absorption bands at 4–5 μm and 13–17 μm, while CH
4 and N
2O also have strong absorption bands at 7.6 μm and 7.8 μm, respectively. However, H
2O (water vapor) is the primary absorber in the entire IR spectral range [
2,
20,
47], with two major absorption bands at 5–8.3 μm and 11–17 μm. The IR flux dominated by CO
2 is well known to be only approximately 10% of that controlled by water vapor [
20]. Therefore, H
2O, together with CO
2, CH
4, N
2O, and halo-GHGs, is the most important GHG, rendering the Earth a unique living environment. However, the atmospheric concentrations of CO
2, CH
4, and N
2O are so high that the atmosphere is completely opaque in the spectral regions of their IR absorption bands. This is clearly seen in the measured and simulated transmittance Γ(
λ) spectrum on the ground using the Sun as a radiation source, showing Γ = 0 in their absorption spectral regions. The OLR transmittance spectrum measured at the TOA by a satellite appears to give a transmittance Γ = ~0.4 in their absorption spectral regions. However, this non-zero transmittance is due to the re-emission of these molecules in the tropopause or stratosphere, as evidenced by the measured OLR flux in the CO
2 absorption band at 14–16 µm corresponding to a blackbody T = ~220 K [
3,
8].
The current debate lies on the question as to whether the absorption at the wings of the IR bands of non-halogen WMGHGs (CO
2, CH
4, and N
2O) would continue to increase with rising gas concentrations and contribute considerably to the observed climate change. For a resolution to this critical question, it is required to have a deep and correct understanding of the basic physics of CO
2 in the atmosphere, as discussed by William Happer [
8]. This key issue is closely related to
water vapor, which has the largest greenhouse effect in the Earth’s atmosphere [
2,
20,
47] and is most likely to be responsible for the largest uncertainties in CO
2 climate models [
2,
17,
20,
22,
47]. As shown in
Figure 1, water has strong and extensive absorption bands overlapping with the extended wings of CO
2, CH
4, and N
2O absorption bands. Interestingly, Kondratiev and Niilisk [
17] made a study of the radiation budget by investigating the absorption and transmission functions of CO
2 and H
2O in the atmosphere in the spectral region of 12–18 μm, showing that the change in the terrestrial radiation with an increased or decreased CO
2 concentration is sharply diminished when the absorption of water vapor is taken into account and that the heat radiation of the atmosphere
is almost independent of the variation in CO2 in the atmosphere. They estimated the ECS to be ≤0.50 K. Similarly Newell and Dopplick [
20] estimated with a static radiative flux model that at low latitudes (the tropics) the increase in surface air temperature due to CO
2 doubling is less than 0.25 K. ECS values of 0.3–0.5 K and 0.26–0.4 K were also observationally determined by Lindzen [
22] and Idso [
21,
23] respectively. These researches showed that the far larger radiative forcings of rising CO
2 and resultant surface T changes computed by GCMs (e.g., the models of Manabe–Wetherald [
48] and Augustsson–Ramanathan [
49]) arise likely from the neglected account of water vapor in the atmosphere and not from the CO
2 itself. Thus, the theory that climatic change is caused primarily by rising CO
2 has been very questionable for decades.
The overlapping effect of water vapor in the spectral ranges of CO
2, CH
4, and N
2O adsorption bands can also potentially cause a
negative feedback factor for the radiative forcings caused by these non-halogen GHGs if the forcings considerably exist. In contrast, water vapor obviously has almost no interference with the radiative forcings of halo-GHGs and ozone due to no or negligible overlapping between the absorptions of H
2O and halo-GHGs or O
3 in the atmospheric window at 8–13 μm. Furthermore, an increase in water vapor in the atmosphere due to surface warming caused by an increase in an effective GHG can lead to a
positive feedback effect [
2,
20,
47], particularly true for halo-GHGs and O
3.
The critical overlapping effect of water vapor on the greenhouse effect of non-halogen WMGHGs (particularly CO2) in the atmosphere must be examined carefully by observations.
3. Brief Review of CO2-Based Climate Models
Climate models aim to estimate the surface T change in response to a change in the atmosphere. A central concept is
radiative forcing, which is a radiative perturbation arising from an increase in a GHS, inducing an initial rise in surface T. This is followed by complex responses including enhanced evaporation of water vapor into the atmosphere, and changes in cloud cover and in the atmospheric or oceanic circulation. In various climate models from the global energy balance models to GCMs, the equilibrium GMST has a linear relationship with the radiative forcing [
1,
2,
13,
50]. Thus, the radiative forcing provides a useful metric to access and compare the impacts of various anthropogenic and natural variations on the Earth’s climate.
Considering the simple radiative balance model of the Earth, one can write the global mean net radiation flux
Fz at the TOA as
Fz = F0 − F↑, where
F0 is the incoming solar flux at the TOA, and
F↑ is the outgoing energy flux from the Earth. For the system to be at equilibrium, the net radiation flux
Fz = 0. When the Earth’s climate system is imposed by a small energy perturbation, for example, by an increase in the atmospheric concentration of a GHG or in incoming solar radiation, this energy imbalance translates into an initial radiative flux perturbation Δ
F at the TOA: Δ
F =
F0 −
F↑, before the surface T changes. This energy flux imbalance is called the radiative forcing (Δ
F). In climate models, however, radiative forcing (
RF) is most commonly computed in terms of the radiative perturbation at the tropopause rather than at the TOA. That is,
F0 and
F↑ after the energy perturbation are retrieved from the model at the tropopause after stratospheric Ts have been allowed to readjust to radiative equilibrium while surface and tropospheric Ts and state variables such as water vapor and cloud cover are fixed at the initial unperturbed values. This definition of radiative forcing (
RF, also called stratospherically-adjusted radiative forcing—
SARF, as distinct from instantaneous radiative forcing) was adopted in the IPCC 2001 TAR [
50], 2007 AR4 [
1] and 2013 AR5 [
2]. Moreover, another measure of radiative forcing, the so-called
effective radiative forcing (
ERF), was introduced in the IPCC AR5 [
2].
ERF is defined as the change in net downward radiative flux at TOA after allowing for atmospheric Ts, water vapor, and clouds to adjust while keeping surface T or a portion of surface conditions unchanged. In GCMs such as CMIP5 and CMIP6, however,
RF and
ERF are practically equal for WMGHGs [
2,
13].
In fact, the calculations of radiative forcing are often simplified into an algebraic formulation specific to a GHG. Since the late 1990s, the IPCC Reports [
1,
2,
50] have used simplified analytical expressions derived from atmospheric radiative transfer models to calculate the
RFs for WMGHGs. Such simplified formulas of
RF are the following [
50,
51]
where
X is the concentration of a halo-GHG in ppb,
χ in Wm
−2 ppb
−1 is the radiative efficiency of a specific halo-GHG, and
C is the concentration of CO
2 in ppm. The subscript 0 denotes the unperturbed concentration (in 1750):
Co = 278 ppm, and
Xo = 0. For CH
4 and N
2O, the radiative forcing is assumed to be approximately proportional to the square root of gas concentration [
50,
51]. Such
RF calculations given by Myhre et al. [
51] have shown excellent agreement (within 5%) with high spectral resolution radiative transfer calculations in GCMs [
50]. Consistently, the author [
3] indeed demonstrated that the
RFs calculated by Equations (2) and (3) are exactly identical to the
ERFs given in the 2013 IPCC AR5 [
2]. As shown in
Figure 2a,b, here we further note that for CO
2, there are small increases of 2–4% in
ERFs computed by GCMs and given in the newest 2021 IPCC AR6 [
13], compared with either
ERFs given in the 2013 IPCC AR5 [
2] or
RFs calculated by Equation (3), whereas for halo-GHGs, there are no differences between
RFs calculated by Equation (2) and
ERFs given in the IPCC AR6 [
13] or AR5 [
2]. Therefore, for simplicity, one can use the simplified functional forms of
RFs, instead of complex GCMs, to access and compare the potential climate effects of various WMGHGs [
3].
Climate change occurs when the climate system responds to counteract the radiation flux changes, and all such responses are explicitly excluded from the definition of
RF in GCMs [
1,
2,
50]. The
climate sensitivity factor (α
c) is defined as the ratio of the (Planck) responsive change in Earth’s surface T (Δ
Ts) to a radiative forcing
RF:
αc is also called the
reference climate sensitivity or
the Planck feedback factor, which is equal to ~0.30 K/(Wm
−2). However, the change (Δ
Ts) in Earth’s equilibrium GMST must also consider the climate feedback factors of water vapor, lapse rate, albedo, clouds, etc. that amplify the (Planck) response to a change in radiative forcing
RF. The change in equilibrium GMST is then given by [
2,
50]
where
β is the
total (feedback) amplification factor and
λc = βαc is called the
equilibrium climate sensitivity factor. Note that in recent IPCC AR5 [
2] and AR6 [
13], there are no separate
αc and
β terms;
λc is instead expressed as the reciprocal of the ‘total feedback parameter’ η, i.e.,
λc = 1/η, where η is the sum of the Planck, water vapor, lapse-rate, cloud, and surface albedo feedback parameters [
52]. Nevertheless, the IPCC AR6 also states that the climate feedbacks amplify the Planck T response by about (
β = ) 2.8 [1.9 to 5.9] times, which was ~3.3 [1.7–5.0] in the AR5, and that cloud feedback contributes to the largest uncertainty of the total feedback. Comparing the observed global surface T rise Δ
Ts = ~0.6 K during the period 1970–2000 with the computed total
RF from all WMGHGs gives rise to a λ
c = ~0.8 K/(Wm
−2) in IPCC’s TAR [
50], which is virtually identical to λ
c = 0.79 K/(Wm
−2) (= 1/η, where η = 1.27 K
−1Wm
−2) obtained from the multi-model mean of GCMs (CMIP3 or IPCC AR4’s coupled atmosphere-ocean GCMs—AOGCMs) [
52].
In CO
2 climate models, however, climate sensitivity is often expressed as the so-called
equilibrium climate sensitivity (
ECS), which is defined as the Δ
Ts (Δ
T×2) that would result from a sustained doubling of atmospheric (equivalent) CO
2. In the IPCC AR6 [
13], climate models have given the best estimate of the ECS to be 3 °C, with a likely range of 2.5–4 °C. This ECS value is identical to that in the IPCC AR4 [
1] and is nearly the same as in the AR5 [
2]. However, it is well-known that the ECS varies largely among climate models, is sensitive to model parameters, and is not well constrained by observations [
26,
33,
34,
53]. Estimates for modern climate conditions give ECS in the wide range of 0.1−10 °C, as reviewed previously [
3,
23,
50]. For the 12 IPCC AR4’s AOGCMs reviewed by Dufresne and Bony [
52], the multi-model mean values are: the net radiative forcing Δ
F is 3.71 ± 0.2 W/m
2 for a doubling of CO
2, λ
c is 0.79 ± 0.19 K/(Wm
−2), and ECS is 3.1 ± 0.7 K. It is noteworthy that these values are virtually equal
to those calculated simply from
RF = 5.35 × ln2 = 3.71 W/m
2 by Equation (3) and Δ
Ts = λcRF = 3.0 K by Equation (5) with λ
c = 0.8 K/(Wm
−2), which are also perfectly identical to the best estimate of ECS = 3 K given in the IPCC AR4 [
1] or AR6 [
13].
Figure 2.
Time-series (effective) radiative forcings and observed and GCM-calculated global mean surface temperature changes. (
a,
b): Time-series (effective) radiative forcings (
RFs/ERFs) of CO
2 and halo-GHGs, relative to the pre-industrial period in 1750, with
RFs calculated by Equations (2) and (3) and
ERFs obtained directly from the IPCC AR5 and AR6 (computed by GCMs), as indicated. Also shown are the linear fits to the
ERFs after 2005 given in the IPCC AR6, with the produced annual rising rates in
ERF indicated. (
c) Observed and calculated global mean surface temperature changes Δ
Ts for 1850–2100, with respect to the mean temperature in 1950–1970. Δ
Ts are calculated by Δ
Ts = λc ×
ERF (Equation (5)) with λ
c = 0.8 K/(Wm
−2) (blue thick line) and transient temperature change Δ
Tst =
λct ×
ERF (Equation (6)) with
λct = 0.51 K/(Wm
−2) (green thick line) for all anthropogenic and natural
ERFs given in the IPCC AR6 for the future projection SSP245. The observed annual GMST data were obtained from the UK Met Office’s HadCRUT5 dataset [
54] (black thin line) and NOAA 5.1 [
55] (red thin line). Updated from Lu [
3] with
ERFs from the IPCC AR6 [
13].
Figure 2.
Time-series (effective) radiative forcings and observed and GCM-calculated global mean surface temperature changes. (
a,
b): Time-series (effective) radiative forcings (
RFs/ERFs) of CO
2 and halo-GHGs, relative to the pre-industrial period in 1750, with
RFs calculated by Equations (2) and (3) and
ERFs obtained directly from the IPCC AR5 and AR6 (computed by GCMs), as indicated. Also shown are the linear fits to the
ERFs after 2005 given in the IPCC AR6, with the produced annual rising rates in
ERF indicated. (
c) Observed and calculated global mean surface temperature changes Δ
Ts for 1850–2100, with respect to the mean temperature in 1950–1970. Δ
Ts are calculated by Δ
Ts = λc ×
ERF (Equation (5)) with λ
c = 0.8 K/(Wm
−2) (blue thick line) and transient temperature change Δ
Tst =
λct ×
ERF (Equation (6)) with
λct = 0.51 K/(Wm
−2) (green thick line) for all anthropogenic and natural
ERFs given in the IPCC AR6 for the future projection SSP245. The observed annual GMST data were obtained from the UK Met Office’s HadCRUT5 dataset [
54] (black thin line) and NOAA 5.1 [
55] (red thin line). Updated from Lu [
3] with
ERFs from the IPCC AR6 [
13].
As shown in
Figure 2c, the calculated Δ
Ts values by Equation (5) arising from the computed total forcing of all anthropogenic and natural ERFs given by the IPCC AR6 are significantly larger than the observed Δ
Ts since 1970. To fit the observed GMST rise during 1970–2002, AOGCMs introduced the so-called
transient T change Δ
Tst [also called the
transient climate response (TCR), instead of equilibrium
T change Δ
Ts], in which an additional parameter, the ocean heat uptake efficiency κ (= 0.69 K
−1Wm
−2) is added to the total feedback parameter η [see, e.g., Dufresne and Bony [
52]]
where
λct = 1/(η + κ) = 0.51 K/(Wm
−2). The thus calculated Δ
Tst results appear to show better agreement with the observed data, as shown in
Figure 2c. Note also that a lag of 10 years for surface (lower-troposphere) measured anthropogenic GHG concentrations and resultant radiative forcings is applied in calculations of or comparisons with surface
T changes throughout this paper [
14,
56,
57], including this Section,
Section 5 and
Section 6. This lag also slightly improves the agreement with observations in
Figure 2c.
At first glance, AOGCMs appear to show excellent agreement with observations, except for cooling magnitudes due to volcanic aerosol eruptions. The GCM-simulated volcanic
ERFs given in the IPCC AR6 are likely largely overestimated when compared with the observed surface temperature record. For example, the well-known Pinatubo eruption produced global cooling of 0.2–0.3 K occurring primarily in 1992 and between 40° S and 70° N, deduced from a robust multivariate analysis of the observed surface temperature record from 1889 to 2006/2008 by Lean and Rind [
56,
57], which is also consistent with the author’s recent analysis of the temperature record from 1850 to 2021 [
14]. In contrast, the volcanic
ERF given in the IPCC AR6 leads to volcanic cooling of 1.43 and 0.91 K in 1992 when λ
c = 0.8 K/(Wm
−2) and
λct = 0.51 K/(Wm
−2) are used respectively, as shown in
Figure 2c. These magnitudes of volcanic cooling are approximately 5 and 3 times larger than the observed ones. Although the volcanic
ERFs, to which the response in surface T occurs relatively rapidly (within months) [
56,
57], have no effect on the simulated long-term trend in GMST, this discrepancy implies that the aerosol
ERFs presented in the IPCC Reports, which are generated from GCM-simulations rather than directly from historical observations, are likely largely overestimated, as further discussed below.
Here we must note that the IPCC-given total radiative forcing
ERF includes large contributions from the
ERF (
RF) of tropospheric (and stratospheric) ozone (
ERFO3) and the aerosol
ERF, which are based entirely and mainly on GCM model simulations, respectively, rather than observational determinations. In the AR6 [
13], the 50% uncertainty (5–95% range) in
ERFO3 remains from AR5 which is largely due to the uncertainty in pre-industrial emissions (though observational ozone data are available since the late 1950s). The AR5 [
2] assessed the total ozone
RF to be +0.35 [0.15 to 0.55] Wm
−2 from 1750 to 2011, which could be split either to be 0.40 [0.20 to 0.60] Wm
−2 for tropospheric ozone and −0.05 ± 0.10 Wm
−2 for stratospheric ozone or to be 0.50 [0.30 to 0.70] Wm
−2 from ozone precursors and −0.15 [−0.3 to 0.0] Wm
−2 from the effect of ODSs. The AR6 gives with high confidence an assessed
ERFO3 of 0.41 and 0.47 [0.24 to 0.70] Wm
−2 over the 1850–2010 and 1750–2019 periods, respectively. These assessments of
ERFO3 are sensitive to the assumptions on precursor emissions used to drive the models. It is summarized in the AR6 [
13] (p. 838): “The CMIP6 model ensemble shows a constant global (tropospheric ozone) increase since the mid-20th century whose rate is consistent with that derived from observations since the mid-1990s”. Thus, there is potentially a larger uncertainty in
ERFO3 for periods prior to the mid-1990s given in the IPCC Reports.
It is a robust fact that the large uncertainty in aerosol
ERF is the dominant contributor to the uncertainty in total
ERF since 1750, given by the IPCC. The estimates of aerosol
ERF are based mainly on GCM model simulations, as reviewed in the IPCC AR5 and AR6. It must also be noted that there are very large changes in the given aerosol
ERF from the AR5 to AR6 (see Table 7.8 of the AR6 [
13]), namely the magnitude of the aerosol
ERF (“ERFari”) due to aerosol-radiation interactions is reduced by about 50% compared to AR5, while the magnitude of the aerosol
ERF due to aerosol-cloud interactions (“ERFaci”) is increased by about 85% compared to AR5, based on the so-called “observation-based and modeling-based evidence”. Although the AR6 assesses that the total aerosol
ERF is virtually certainly negative, large uncertainty remains, particularly regarding the adjustment contribution of aerosol-cloud interactions to
ERF, as well as missing processes in current Earth system models, notably aerosol effects on mixed-phase, ice, and convective clouds. Note also that observational constraints (e.g., by the observed temperature record) on the total aerosol
ERF are made often by combination with GCMs, assuming that the historical evolution of the total forcing of GHGs would be well constrained [
13]. If this assumption in GCMs, particularly a dominantly large
ERF from CO
2, is seriously incorrect (as unfortunately demonstrated by observational fingerprints in
Section 5), then all such estimates are invalid. Even so, the estimates based on energy balance considerations or other observational constrains are strongly model-dependent and give an extremely wide range of −2.0 to 0 Wm
−2 for the total aerosol
ERF (see Figure 7.5 of the AR6 [
13]). Thus, it is clear that there are very large uncertainties in
ERFO3 and aerosol
ERF given by GCMs and IPCC Reports.
For an analysis, we show measured aerosol loadings (sulfate and black carbon—BC) from ice-core measurements, measured ozone climatology from the WOUDC’s TOST ozonesonde datasets, and GCM-simulated
ERFs in
Figure 3, all of which are used or presented in the IPCC AR6 [
13]. As stated in the AR6, data from ice cores allow for the estimation of multi-centennial trends of aerosol deposition in mid- and high-latitudes, as plotted in
Figure 3a,b. In these regions, sulfate or BC concentrations exhibited large increases by factors of 2–8 from the end of the 19th century to the 1970s, namely reaching their peak around 1950 in the Arctic and 1970 in Europe and Russia, and have declined by about a factor of about 2 thereafter. South America has kept a small positive trend in either sulfate or BC concentration since 1750. Of particular interest are the sulfate and BC concentrations in the Antarctic, which have exhibited extremely low levels and no increases since 1750, and in the Arctic, which have returned to their pre-industrial levels since around 2000. According to the AR6, these changes in aerosol concentration are also consistent with and are likely the main contributor to observed changes in Earth’s surface energy budget characterized by surface solar radiation, which exhibits that downward and upward thermal radiation has increased since the 1970s. Satellite- and ground-based data indicate that aerosol optical depth (AOD) exhibits predominantly negative trends since 2000 over Northern Hemisphere (NH) mid-latitudes and Southern Hemisphere (SH) continents while increasing over South Asia and East Africa. These lead to a globally deceasing aerosol abundance assessed with medium confidence.
As shown in
Figure 3c, the observed TOST ozone data for the period 1960s–2000s [
38] are practically consistent with the conclusion in the IPCC AR6 [
13] (p. 307), which states that tropospheric ozone increased by 30–70% across the NH from the mid-20th century to the mid-1990s and that since then, free tropospheric ozone has increased by 2–7% per decade in most regions of NH mid-latitudes, 2–12% per decade in the sampled regions of the NH and SH tropics, and less than 5% per decade at SH mid-latitudes. Note that the TOST datasets used in
Figure 3c are also used in the AR6.
As shown in
Figure 3d,e, the above-measured data of aerosol loadings and ozone are somewhat reflected in
ERFs simulated by GCMs (CMIP6) and given in the AR6, though there are large uncertainties in modeled
ERF magnitude, especially for periods prior to the mid-1990s. For example,
Figure 3e shows that the aerosol ERF changed largely and negatively from about zero in 1850 to −1.4 Wm
−2 in ~1975, became almost constant during 1975–2010, and has become less negative thereafter.
Figure 3f interestingly shows that the sum of O
3 ERF and aerosol
ERF exhibits a large negative net
ERF of about −1.1 Wm
−2 from 1850 to ~1970, followed by
a very small change (within 0.1 Wm−2) between 1970 and 2010. It is noteworthy that under the robust observational constrain of a very small rise in regional or global surface temperature of about 0.2 K after removal of natural effects from 1750 (1850) to 1970, this large negative net
ERF (mainly aerosol
ERF) for the period 1850(1750)–1970 is absolutely required to offset the large positive modeled
ERF arising from CO
2 (CH
4 and N
2O) since the Industrial Revolution, which is +1.30 to +1.45 Wm
−2 in 1970–1975, in CO
2-based GCMs.
The AR6 concludes that the total
ERF due to all anthropogenic and natural climate drivers is positive and has grown in magnitude since the late 19th century and that the change rate has likely increased in the last 3 decades due to the observed increasing growth rate of CO
2 and the changing to become less negative in aerosol ERF (
Figure 3e). However, it is worth noting that
Figure 3f also indicates that the net of ozone and aerosol ERFs made very little contribution to the observed rapid surface warming during 1975–2005 (the late 20th century), while both climate drivers (ozone and aerosols) are well known to cause a rather quick response in surface temperature (on timescales of months only), as well observed from the Antarctic ozone hole and past volcanic eruptions.
There are other contributors to the total
ERF, including stratospheric water vapor, land use, surface albedo (black plus organic carbon aerosol on snow and ice), combined contrails, and aviation-induced cirrus [
13]. However, these contributors are relatively minor, far smaller than the uncertainties of
ERFO3 and the aerosol
ERF. Thus, they will not be considered in this study.
Note also that it is the calculated Δ
Ts by Equation (5) with λ
c = 0.8 K/(Wm
−2), rather than the calculated Δ
Tst by Equation (6) with
λct = 0.51 K/(Wm
−2), for a doubling of CO
2 (
RF = ~3.71 W/m
2), that matches perfectly the best estimate of ECS = 3 K in the IPCC AR4 [
1] or AR6 [
13]. The calculated Δ
Tst (=1.89 K) for a CO
2 doubling is obviously much lower than the best-estimated ECS. One may argue that Δ
Tst in AOGCMs is not an equilibrium
T change that might take a multi-century to millennial timescale to reach. However, the natural climate drivers such as 11-year solar cyclic variability, ENSO, volcanic eruptions, as well as sudden surface temperature changes in the Antarctic due to ozone depletion since the 1960s (to be shown in
Section 5), have taught that the observed surface temperature response to natural radiative forcing occurs rather rapidly (within months) [
3,
14,
56,
57,
58]. Thus, it seems quite reasonable to assume that the measured GMST with a lag of 10 years from anthropogenic
RF (
ERF) is approximately the equilibrium
T [
14,
56,
57]. In any case, a major aspect in introducing the ocean heat uptake efficiency as an additional parameter to GCMs is apparently to better match the models to the measured GMSTs, as shown in
Figure 2c. One can see from
Figure 2c that Equations (5) and (6) serve well to show the prediction capability of GCMs or AOGCMs and their calculated results of long-term GMST changes.
According to the IPCC AR5 [
2] and AR6 [
13], there are major differences in definition between
RF and
ERF. In contrast to
RF,
ERF is defined as the change in net downward radiative flux at TOA after allowing for atmospheric Ts, water vapor, and clouds to adjust in both the stratosphere and troposphere. If the modeled results on the overlapping effect of water vapor on the climate forcings of CO
2 in the literature [
17,
18,
19,
20] are correct, which will be tested with the key observations reviewed in
Section 5, it will be quite surprising to see no significant differences between
RF and
ERF or even a small
increase rather than decrease in
ERF for CO
2 (the tropospheric water vapor is increased when the surface is warmer). This may imply that there exist severe problems in the fundamental assumption of no significant overlapping (nullifying) effect of water vapor on the CO
2-produced radiative forcing in CO
2 climate models.
4. Early Calculations of Global Warming Caused by Halo-GHGs
In 1975, Ramanathan [
59] made the first calculations that the greenhouse effect by CFCs and chlorocarbons could lead to a rise of ~0.9 K in global surface T
if each atmospheric concentration of these compounds would increase to 2000 ppt. This projected concentration is far larger than the observed values due to the successful Montreal Protocol. Subsequently, Ramanathan and co-workers [
49,
60,
61,
62] and Wang and Molnar [
63] also studied the climatic effects of CO
2 and non-CO
2 GHGs (CFCs, CH
4, N
2O, O
3, and others), showing that CFCs, through their indirect O
3-depleting effect, would have a potentially large stratospheric cooling effect, as large as that due to the CO
2 increase. Fisher et al. [
64] also calculated the greenhouse effect of HCFCs. Ramanathan [
65] concluded that the non-CO
2 GHGs contributed to the radiative heating of the surface-atmosphere system as much as CO
2 during the middle to late 20th century.
Interestingly, Wang et al. [
66,
67] showed that the spatial distribution of atmospheric opacity which absorbs and emits the long-wave radiation for halo-GHGs (CFCs) is different from that for CO
2. Their simulations indicated that halo-GHGs provide an important radiative energy source for the Earth climate system and that different infrared opacities of CO
2 and halo-GHGs can lead to different climatic effects. They concluded that it is inappropriate to use an ‘effective’ CO
2 concentration to simulate the total greenhouse effect of CO
2, CFCs, and other GHGs. Forster and Joshi [
68] used various climate models to examine the role of halo-GHGs on stratospheric and tropospheric Ts. They found that halo-GHGs (mainly CFC-12) would have contributed a significant warming of ~0.4 K at the tropical tropopause since 1950, dominating the effect of other WMGHGs. They also noted that the “disappearance” of such T increases would suggest that some other mechanism(s) such as stratospheric cooling due to O
3 loss are highly likely to be compensating for this, and as O
3 will likely recover in the next few decades, a slightly faster rate of warming would be expected from the net effect of halo-GHGs.
Since the 2000s, Ramanathan and co-workers [
69,
70,
71,
72] have turned to emphasize the importance of black carbon (BC) and atmospheric brown clouds in causing regional and global warming. They estimated that when globally averaged, the current BC radiative forcing at the TOA is as much as 60% of the current
RF due to the greenhouse effect of CO
2, and therefore, concluded that BC is potentially the second major contributor to the observed 20th-century global warming, just next to CO
2. Thus, halo-GHGs would become the third contributor to the warming.
Today, it is well known that CFCs are much more potent GHGs than CO
2 on a molecule-by-molecule basis, and the radiative impacts of halo-GHGs on global climate change have received gradually increased attention [
62,
73,
74,
75]. However, the contribution of halo-GHGs has only been assessed under CO
2 climate models, and the newest IPCC AR6 [
13] gives that CO
2 and halo-GHGs have currently accounted for ~65% and 12%, respectively, of the GCM-computed total
ERF from all WMGHGs with respect to their pre-industrial concentrations in 1750 and for ~82% and ~4%, respectively, for the
ERF increase in the past decade. All these studies unavoidably concluded that CO
2, rather than halo-GHGs, would make the primary contribution to climate change, and that the total
ERF (GMST) rising rate has increased in the last 30 years due to the observed increasing growth rate of CO
2 and the changing to become less negative in aerosol
ERF [
2,
4,
13] (
Figure 2 and
Figure 3). However, the assumed large greenhouse (forcing) effect of non-halogen WMGHGs under CO
2 climate models (GCMs) is yet to be validated by a series of critical observations below.
6. Refined Physical Model of Global Climate Change and Parameter-Free Theoretical Calculations of GMST Changes through the CFC-Warming Model
Let us briefly revisit the fundamental concept of radiative forcing. A small concentration increment of a GHG initially
increases the absorption of OLR in the atmosphere in the wavelength range of the GHG IR absorption, which leads to an initial
decrease in OLR at the TOA. To compensate for this decrease in OLR at the TOA, the Earth’s surface must raise its
T to make an increase (d
Fg) in upward long-wave radiation energy flux
Fg. That is, the increase in
Fg from the warming must balance the reduction in OLR at TOA due to the increased absorption of the GHG. Finally, a new equilibrium is restored at TOA. This is the well-known basic mechanism underlying global warming. According to this classical mechanism and the quantum physics of the Earth’s blackbody radiation (Equation (1) and
Figure 1), the radiative forcing due to a GHG rise can be written as
where
Bλ(
Ts) is given by Planck’s formula (Equation (1)), and Γ(
λ) is the wavelength-dependent transmittance of the atmosphere (
Figure 1). As discussed in
Section 2, a change in absorption in the atmospheric window sensitively influences the radiative process of the Earth. Since most of the Earth’s thermal radiation is emitted into space through the atmospheric window, Equation (7) can approximately be rewritten as
where Γ
wd is the (mean) transmittance in the atmospheric window at 8–13 µm, and
σ is the Stefan-Boltzmann constant. Note that Γ
wd is significantly different from the average transmittance Γ over the entire blackbody radiation spectrum. Hence, the climate sensitivity factor α is now given by
Here Γwd can be determined from the OLR spectrum at the TOA measured by a satellite. Various infrared opacities of GHGs in the atmosphere can lead to different climatic effects. Below we will discuss the greenhouse effects of anthropogenic GHGs in the two major categories: non-halogen GHGs and halo-GHGs.
Non-halogen GHGs (CO2, CH4 and N2O). In view of the key observations revealing no warming effects (forcings) of these non-halogen GHGs (
Figure 4,
Figure 5,
Figure 6,
Figure 7,
Figure 8 and
Figure 9), we may as well let the
RF due to an increase in CO
2 (CH
4 or N
2O) since 1970 be zero (
RF = 0). Alternatively, one might attribute the nullifying effect of water vapor on the IR absorptions of these non-halogen GHGs to the feedback amplification factor and let the latter be zero (β = 0). In either case, this leads to virtually no change (Δ
Ts = 0) in GMST caused by a concentration rise of non-halogen GHGs (CO
2, CH
4, and N
2O) since 1970. This conclusion can also be deduced either from the modeled results of the nullifying (overlapping) effect of water vapor on the climate forcings of CO
2 or from the observational determinations of the ECS, both of which gave the ECS = 0.25–0.5 K [
17,
18,
19,
20,
21,
22,
23,
26]. Given the radiative forcing
RF = 3.7 W/m
2 for a doubling of CO
2 by Equation (3) or given in the IPCC AR5/AR6, then the forcing
RF = 0.7 W/m
2 arising from the CO
2 increase in the rapidly warming period 1970–2002 should lead to an increase of 0.05–0.1 K only in GMST. Note that if one argues for any delay in surface T response to CO
2 RF due to ocean heat uptake, then this temperature increase would even become smaller because the CO
2 rising rate has been increasing since 1750. This small surface T change is within the general uncertainty of measurements for GMST and is therefore negligible (Δ
Ts ≈ 0).
Halo-GHGs such as CFCs and HCFCs. From the satellite-measured atmospheric transmittance Γ(
λ) spectrum shown in
Figure 1, we can directly obtain the mean transmittance Γ
wd = 0.84 in the window of 8–13 µm. Substituting this measured Γ
wd and
Ts = 288 K into Equation (9), we obtain a climate sensitivity factor for halo-GHGs [
3]
The total amplification factor β is well-known to have the largest uncertainty due to the feedback effects of water vapor and clouds in GCMs [
2,
13,
47]. To determine the value of β more reliably, in contrast, we took an observational approach by a careful analysis of well-observed GMST variations arising from solar irradiance variability during 11-year solar cycles [
3]. First, a solar radiative forcing Δ
F0 arising from a change in the incoming solar energy flux
F0 at the TOA in solar cycles can reliably be calculated from the measured change in solar constant (
FS) (also called total solar irradiance, TSI) by
where
A is the albedo of the Earth-atmosphere system (
A ≈ 0.3). Second, the direct (Planck) response in surface
T to the solar radiative forcing is simply given by
From Equations (11) and (12), we can write the solar climate sensitivity factor (
αS)
With the observed values of
TS = ~288 K, Δ
FS/
FS = ~0.1% from solar minimum to solar maximum in solar cycles, and
FS = ~1361 Wm
−2, we obtain Δ
TS =~0.072 K and α
c = 0.30 K/(Wm
−2). Since the variation of GMST in solar cycles is well observed to about 0.11 K with a rapid response time of only one month to solar forcing [
56,
94], we directly obtain the total feedback amplification factor β = 1.53 [
3].
The thus determined solar equilibrium climate sensitivity factor is [
3]
which is slightly smaller than
λcs = 0.63 K/(Wm
−2) obtained by Douglass and Clader [
58] with a similar approach. But they used a slightly larger temperature modulation of 0.15 K instead of 0.11 K during solar cycles. Thus, our λ
cs value should be more realistic.
With
αhalo = 1.16 K/(Wm
−2) and
β = 1.53, we explicitly obtained the equilibrium climate sensitivity factor for halo-GHGs [
3]
Combining the greenhouse effect of halo-GHGs with the solar radiative effect, we obtained the change (Δ
Ts) in GMST as a result of a rise in halo-GHG concentration and a change in solar radiative output [
3]
where the
RFhalo of halo-GHGs can be calculated by Equation (2), and
RFsolar is calculated from the measured solar constant (
FS or TSI) by Equation (11) for historical data and a simplification λ
cs ×
RFsolar ≈ 0.055 × cos [2π (i-2014)/11] (°C) is used for future projections, based on the observed temperature modulation during solar cycles [
56,
94] (for details, see ref. [
3]).
Thus, our conceptual physical model of climate change, built on the key observations reviewed in
Section 5, includes
no tunable parameter. It needs only inputs of atmospheric concentrations of halo-GHGs, which are available from the IPCC AR5 [
2] or AR6 [
13]. Using this simplified physical climate model (Equation (16)), the author [
3,
14] has calculated GMSTs from 1950 to 2070 and compared them with the observed data with and without the the removal of natural ENSO and volcanic effects since 1950. As shown in
Figure 10, the calculated results of GMST exhibit surprisingly good agreement with the observed GMST data since the 1950s [
14]. Remarkably, the calculated results also match nearly perfectly with the observed GMST data since 2015 if the regions in Russia and Alaska are excluded [
14]. These results strongly indicate that
the change in GMST since 1950 has been dominantly caused by the greenhouse effect of halo-GHGs (mainly CFCs).For comparison, the calculated Δ
Tst results induced by all natural and anthropogenic
ERFs using the GCM-computed
ERFs given in the IPCC AR6 and Equation (6) with the
λct value given in AOGCMs [
1,
2,
50] are also shown in
Figure 10. It is clearly shown that the fundamental assumption on the climate forcing of increasing non-halogen GHGs inevitably leads to a continued warming trend over the past two decades and in the rest of this century, which is inconsistent with the crucial observations reviewed in
Section 5.
Now we utilize our CFC-warming physical model to analyze the surface temperature change in the Antarctic shown in
Figure 9c, which is caused primarily by the greenhouse effect of halo-GHGs and ozone depletion caused by ODSs. Note that like halo-GHGs, the IR absorption of ozone at 9.6 μm lies in the atmospheric window (
Figure 1). Thus, λ
chalo is also valid for quantifying the greenhouse effect of O
3. First, the
ERF arising from the increase in total halo-GHG concentration is 0.177 Wm
−2 from 1960 to 1980 and 0.171 Wm
−2 from 1980 to 2000. Since the observed Antarctic surface T is nearly constant in 1980–2000 (
Figure 9c), the latter
ERF value must be balanced by the negative
ERF caused by ozone depletion in the Antarctic troposphere and stratosphere. As a result, we deduce that
ERFO3 = −0.171 W/m
2 for the period 1980–2000. The measured decrease in total ozone in the period 1960–1980 is about 30% of the decrease in 1980–2000, as shown in
Figure S5 in SI. Thus, the
ERFO3 = γ × (−0.171) × 30% = −0.051γ (Wm
−2) for the period 1960–1980, where γ = 0−1 taking into account the non-linearity of the radiative efficiency of ozone [
95] (γ = 0 represents a complete saturation for the undisturbed O
3 concentration with no significant ODSs in the atmosphere, and γ = 1 means no saturation, i.e., a linear dependence of
ERFO3 on ozone concentration). Since the warming ΔT
S = 0.27 K is observed during 1960–1980, we obtain λ
c = 0.27/(
ERFhalo +
ERFO3) = 0.27/(0.177−0.051γ) (K/Wm
−2). Thus, we obtain λ
c(min) = 1.52 K/(Wm
−2) for γ = 0 (very unlikely, as seen from
Figure 1), and λ
c(max) = 2.14 K/(Wm
−2) for γ = 1. The realistic γ value is known to be 0 < γ < 1 [
95]. It turns out that λ
chalo = 1.77 K/(Wm
−2) in Equation (15), which was obtained previously by our observational determination from 11-year solar variability [
3], perfectly centers at this λ
c range estimated from the observed surface T change in the Antarctic. The individual contributions of halo-GHGs and ozone depletion and the net change to the Antarctic surface T in 1960–1980, as calculated by our CFC-warming model, are shown in
Figure 9d, exhibiting perfect agreement with the observation. This agreement additionally gives strong evidence of our CFC warming model and validates the value of λ
chalo determined from our observational approach [
3].
The good agreements between the observed and calculated results in
Figure 9d and
Figure 10 are the most compelling message delivered by the CFC-warming model. However, we have to discuss the uncertainty in our calculations of the change in GMST since 1950 (1970), arising potentially from the inclusion of only the greenhouse effect of halo-GHGs and the solar forcing in Equation (16). We have not included the
RFs (
ERFs) of ozone, aerosols, and other relatively minor climate drivers in our calculations of the GMST change in the late half of the 20th century, especially since 1970, for the following reasons. First, there have been much more abundant observation data available after 1970, which could potentially better constrain the modeled results of
RFs (
ERFs) than in pre-1970 periods. Second, the net
ERF of O
3 and aerosol
ERFs exhibits a very small change (within 0.1 Wm
−2) between 1970 and 2000s (
Figure 3f), while our main goal is to unravel the primary underlying mechanism for the rapid warming that occurred from around 1975 to 2000. For comparison, we note again that an
ERF as large as 1.30–1.45 Wm
−2 from non-halogen WMGHGs alone or a large negative net of about −1.1 Wm
−2 of O
3 and aerosol
ERFs from 1850 to ~1970 has been given by the IPCC AR6 (GCM simulations), whereas only a very small GMST rise of ~0.2 K excluding natural effects but including the AA effect was observed during this period. Thus, such a small net
ERF of ≤0.1 Wm
−2 should lead to a negligible contribution to the GMST change. Third, we believe that there are very large uncertainties in simulated
ERFs given in the IPCC AR6, and new research beyond our current efforts will be needed to calculate more reliable
ERFs of aerosols, ozone, and other minor drivers. For an approximation, we estimate that the uncertainty in our calculated GMSTs from 1950 to the present is approximately 0.1 °C, which is about 20% of the temperature change. This uncertainty is larger than the estimated GMST rise due to the CO
2 increase in the most rapid warming period 1970–2000 either from the climate models taking into account the overlapping effect of water vapor or from the observational determinations of the ECS [
17,
18,
19,
20,
21,
22,
23]. In other words, the GMST change due to the variations of non-halogen WMGHGs (CO
2), ozone, and aerosols in the late 20th century is negligibly small and is therefore not included in our CFC-warming model calculations via Equation (16). Note also that although the AA effect and polar ozone depletion obviously play an important or dominant role in changing the surface temperatures of the Arctic and the Antarctic, respectively (as seen in
Figure 9b,c), each area of the Arctic and Antarctic regions only takes 6.7% of the global surface area and therefore the AA effect or polar ozone depletion makes only a small contribution to the change in GMST. Therefore, it is generally agreed that uncertainty in GMST change under GHG forcing is dominated by water vapor and cloud feedbacks only, rather than by the AA effect or polar ozone depletion [
2,
13]. Our calculated results clearly demonstrate that the greenhouse effect of halo-GHGs alone has well accounted for the observed GMST data since the 1950s, without the necessity of introducing the ocean heat uptake efficiency parameter κ as in AOGCMs. This indicates that the response in GMST to anthropogenic
RF is relatively rapid (within 10 years, including the lag of mixing and transports of WMGHGs in air circulation on the global scale) [
14,
56,
57]. In view of the key observations reviewed in
Section 5, there is a strong rationale to predict that a long-term cooling trend will occur in the coming decades.
7. Remarks and Perspectives
This paper identifies and tests the critical assumption of a large climate forcing of rising CO2 in traditional climate models by the six key observations from ground-based and satellite measurements. These observations provide fingerprints of the contribution of CFCs versus CO2 to global warming in the late 20th century. It is clearly shown that CO2-based climate models (GCMs) do not pass the validation by any of the key observations. The observations strongly indicate that global climate change has been primarily caused by halo-GHGs since ~1970, with some disguises by ozone recovery and Arctic amplification of surface warming due to continued (sea) ice loss in high-latitude (polar) regions. CO2 climate models show large discrepancies with the six critical observations, arising likely from the neglecting of the overlapping effect of water vapor on the climate forcing of CO2 (N2O and CH4). A critical review of the observed surface temperature changes in the three representative regions (central England, the Arctic, and the Antarctic) shows that the surface temperature changes are primarily caused by the greenhouse effect of halo-GHGs, rather than by the observed increasing growth rate of CO2 and the change in aerosol ERF, though the climate forcing of Antarctic ozone depletion and the AA effect also make large contributions to the surface temperature changes in the Antarctic and Arctic, respectively. These observations agree well with the CFC-warming physical model, whereas they are in marked contrast to the main conclusion in the IPCC AR6. The observations also show that a reversal in global warming would have occurred since around 2005 if there were no Arctic (polar) amplification. Remarkably, the analytically calculated surface temperature changes by the warming physical model of halo-GHGs with no tunable parameter, built on the quantum physics of the Earth’s blackbody radiation and the observational determination of the (equilibrium) climate sensitivity factor, exhibit excellent or perfect agreement with the observations of GMST and particularly the surface temperature in the Antarctic. Overwhelmingly the observed and calculated results support an emerging picture that halo-GHGs have played a dominant role in causing global warming in the late 20th century.
Since GMST is still around the peak over the past 150 years, sea-ice melting in the Arctic unavoidably remains and, in the Antarctic, may start to increase significantly with the emerging recovery of the Antarctic ozone hole (due to the increased greenhouse effect of recovering ozone). Thus, the polar amplification of surface warming due to sea-ice loss at certain Arctic or Antarctic coastal regions may continue to be significant for some years, and therefore regional warming in those polar areas may last for one to two decades until sea ice is stabilized, even if GMST has been stabilized or reversed.
Ice melting on land (glaciers, small ice caps, and ice sheets) and ocean water expansion due to warming can contribute to global mean sea level (GMSL) rise [
13,
96,
97]. However, this rise may be a complex of natural and anthropogenic drivers. The rise in GMSL began about 20,000 years ago when the last glacial phase of the Pleistocene ice age ended and the continental ice sheets began to melt and continued for the ensuing 12 millennia until reaching a GMSL level a few meters below the present level ~7000 years ago [
98]. For the period 8000–1000 years ago, GMSL rose much more gradually. Changes in sea level have been dominated more by regional and global than by glacial melt in the past millennium [
98]. At present, GMSL rises are caused largely by the thermal expansion of seawater and by the calving of Antarctic and Greenland ice sheets. The IPCC AR6 [
13] concludes with high confidence that GMSL is rising, and the GMSL rise rate since the 20th century has been the largest over the previous three millennia. However, recent marine oxygen-isotope-based GMSL reconstructions reviewed in the AR6 also show that GMSL during the Holocene was among the highest (5–25 m higher than the present level), and was surpassed only during the LIG (5–10 m higher) and MIS 11 (−3.5 to +0.5 m different) over the past 800 kyr. A (US) National Academies of Sciences, Engineering, and Medicine Report on Climate Stabilization Targets in 2011 [
96] stated that GMSL from satellite measurements was rising at 3.1 ± 0.4 mm y
−1 since the records began in 1993 to 2003, but this rate decreased somewhat in the period 2003–2008 to 2.5 ± 0.4 mm y
−1 due to a reduction in ocean thermal expansion, whereas contributions from glaciers, small ice caps, and ice sheets increased. Moreover, some analyses based on observed trends rather than model results showed that at the 95% confidence level, no consistent or substantial evidence exists that recent rates of rise are higher or abnormal in the context of the historical records available for the data-rich United States [
98,
99]. Because oceans respond slowly to global warming and hence GMSL rises lag from Earth T changes, the rise in GMSL is expected to continue for coming centuries as a response to the past warming (even if GMST is stabilized through the phasing out of GHGs) [
96,
97]. Thus, there seems no conclusive evidence from recent trends of GMSL that global warming is continuing or stopping or is even caused by a main anthropogenic or natural driver.
This study leads to an important perspective, which may be regarded as good news, that with the phasing out of halo-GHGs (CFCs, HCFCs, HFCs, and PFCs) by international Agreements, including the most successful and important Montreal Protocol and its Amendments, it is very likely to see a gradual reversal in GMST in the coming decades. Nevertheless, this expected reversal in global warming and the emerging shrinking of the ozone holes, including the recently discovered tropical ozone hole that affects approximately half of the world’s population [
16,
38,
100], will come true only with continued international efforts in phasing out all halogenated ODSs and halo-GHGs. Therefore, this study highlights the importance of such efforts from international governments and the global community. However, it is equally important to emphasize that the relevant international policies and political agenda must be built on a solid scientific foundation that has a correct understanding of the underlying science of global climate change. The 2021 Nobel Prize Committee in Physics gave a correct attribution of the primary cause of warming to being anthropogenic, yet to identify a correct GHG species as the primary anthropogenic driver is of critical importance for humans not only to reverse the climate change but to maintain a healthy economy and ecosystem around the globe [
29].
This study and the recent advance in quantitative understanding of global ozone depletion through the CRE theory [
16] lead to the following implications. Geoengineering has been proposed as a potential method to reduce climate warming by increasing sunlight reflection through the intentional addition of aerosols into the stratosphere, known as ‘stratospheric aerosol injection’ [
44]. In view of both the observed evidence of the reversal in climate change since ~2005 and the significant enhancement (harmful) effect of atmospheric cloud and/or aerosol particles on global ozone depletion [
16], this study strongly recommends the proposed geoengineering project
not be proceeded.
On GCMs, some literature reviewed by Balaji et al. [
31] has argued that the entire project of parameterization, i.e., the discovery of parsimonious representation through insight or mathematical methods, may have no future for GCMs, and that large-scale computation may be needed. For a future generation of GCMs, Balaji et al. have proposed to overcome the shortcomings of current GCMs through substantially higher resolution and detail, or through the use of machine learning techniques to match them better to observations, theory, and process models. They anticipate that with the new methods, understanding of climate processes would lie in an entire hierarchy of models where GCMs would continue to play a central role for foreseeable future climate modeling. Here we argue that the assumptions in any climate model including GCMs must carefully be validated with observations, however. Given the six key observations reviewed in this study, the fundamental assumption in GCMs seems invalid. Thus, it is very likely for GCMs to fail in making a correct prediction for long-term climate change whatever how complex and sophisticated they are. In contrast, it may be optimistic to anticipate that the presented conceptual physical model of climate change and insights provided by this study will help to reach proper and right decisions for humans to solve the important climate problem.