Longitudinal population-level HIV epidemiologic and genomic surveillance highlights growing gender disparity of HIV transmission in Uganda ========================================================================================================================================== * Mélodie Monod * Andrea Brizzi * Ronald M Galiwango * Robert Ssekubugu * Yu Chen * Xiaoyue Xi * Edward Nelson Kankaka * Victor Ssempijja * Lucie Abeler Dörner * Adam Akullian * Alexandra Blenkinsop * David Bonsall * Larry W Chang * Shozen Dan * Christophe Fraser * Tanya Golubchik * Ronald H Gray * Matthew Hall * Jade C Jackson * Godfrey Kigozi * Oliver Laeyendecker * Lisa A. Mills * Thomas C Quinn * Steven J. Reynolds * John Santelli * Nelson K. Sewankambo * Simon EF Spencer * Joseph Ssekasanvu * Laura Thomson * Maria J Wawer * David Serwadda * Peter Godfrey-Faussett * Joseph Kagaayi * M Kate Grabowski * Oliver Ratmann * Rakai Health Sciences Program * PANGEA-HIV consortium ## Abstract HIV incidence in eastern and southern Africa has historically been concentrated among girls and women aged 15-24 years. As new cases decline with HIV interventions, population-level infection dynamics may shift by age and gender. Here, we integrated population-based surveillance of 38,749 participants in the Rakai Community Cohort Study and longitudinal deep sequence viral phylogenetics to assess how HIV incidence and population groups driving transmission have changed from 2003 to 2018 in Uganda. We observed 1,117 individuals in the incidence cohort and 1,978 individuals in the transmission cohort. HIV viral suppression increased more rapidly in women than men, however incidence declined more slowly in women than men. We found that age-specific transmission flows shifted, while HIV transmission to girls and women (aged 15-24 years) from older men declined by about one third, transmission to women (aged 25-34 years) from men that were 0-6 years older increased by half in 2003 to 2018. Based on changes in transmission flows, we estimated that closing the gender gap in viral suppression could have reduced HIV incidence in women by half in 2018. This study suggests that HIV programs to increase HIV suppression in men are critical to reduce incidence in women, close gender gaps in infection burden and improve men’s health in Africa. ## Introduction Despite the widespread availability of HIV prevention and treatment interventions, there were 1.5 million new HIV infections and 680,000 HIV-associated deaths in 20201. More than half of these new cases and deaths were concentrated in the eastern and southern regions of the African continent, where incidence rates have historically been highest in adolescent girls and young women, aged 15-24 years 2,3,4,5. While HIV incidence has declined by 43% in eastern and southern Africa since 2010, current HIV service programs are failing to reduce new cases rapidly enough to meet United Nations health targets for HIV epidemic control 1. With rising levels of HIV drug resistance 6,7 and flatlined global investment in HIV control 8, the African HIV epidemic has reached a critical inflection point 9. Over the last decade, African HIV control programs, including the United States President’s Emergency Plan for AIDS Relief (PEPFAR), have focused on expanding treatment coverage in people with HIV and reducing HIV infections among adolescent girls and young women 10,11. However, recent data from Africa indicate that the mean age of infection is shifting 12,13 and incidence rates are declining faster in men than in women 14,15, suggesting that the age and gender structure of the African HIV epidemic is evolving. Here, we integrate 15 years of data on HIV incidence and onward transmission to show how the drivers of the heterosexual African HIV epidemic are changing by age and gender. We focus on a study population aged 15 to 49 years with an HIV risk profile typical across eastern and southern Africa 16,17, living in 36 semi-urban and rural agrarian communities that are part of the population-based Rakai Community Cohort Study (RCCS) in south-central Uganda 18 (Fig. 1a). We followed individuals in the RCCS who were HIV seronegative and documented new infection events. We also deep-sequenced HIV virus longitudinally from persons with viremic HIV. This enabled us to infer directed transmission networks across age and gender 19,20, and focus on the time trends in infection dynamics and transmission networks during mass scale-up of HIV services in Africa 1. ![Fig. 1:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F1.medium.gif) [Fig. 1:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F1) Fig. 1: Time trends in age-specific HIV incidence rates for men and women in Rakai, Uganda. (**a**) Location of the Rakai Community Cohort Study (RCCS) in south-central Uganda. Study outcomes are reported for all RCCS communities located inland to Lake Victoria across nine survey rounds. Sources of the map data: OpenStreetMap open data by OpenStreetMap contributors, see [https://www.openstreetmap.org/copyright](https://www.openstreetmap.org/copyright). (**b**) Number of RCCS participants in the censuseligible population of age 15 to 49 by survey round. (**c**) Estimated mean HIV incidence rates per 100 person-years of exposure in uninfected individuals (line) by 1-year age band, gender and survey round, along with 95% confidence intervals (ribbon), and median age of incident cases (cross). (**d**) Estimated median contribution to incidence cases in the study population (line) by 1-year age band, gender and survey round, along with 95% confidence intervals (ribbon). Throughout all subfigures, incidence estimates are based on *n* =1,117 individuals in the incidence cohort. ## Results ### HIV incidence is declining faster in men than women From September 23, 2003 to May 22, 2018, 38,749 participants were enrolled in the Rakai Community Cohort Study 14. Of these participants, 22,724 tested HIV seronegative at first survey, and contributed an estimated 127,217 person-years of follow-up (Fig. 1b, Supplementary Tables S1-S2). Study participants were enrolled following population census, household enumeration, and informed consent in 9 survey rounds of approximately 18 months duration, herein denoted as survey rounds 10-18 (see Methods and Extended Data Fig. 1). In total, we observed 1,117 incident HIV infections (Supplementary Tables S3-S4 and Extended Data Fig. 2). Fig. 1c shows that incidence rates among men in inland communities fell rapidly from 1.05 [1.03-1.08] per 100 person-years (PY) in 2003 (survey round 10) by 67.8% [66.2-69.2] to 0.34 [0.33-0.35] per 100 PY in 2018 (survey round 18), with no substantial shift in the median age of male incident infection (blue triangles in Fig. 1c). In young women aged 15 to 24 years, incidence rates fell similarly rapidly from 1.42 [1.35-1.5] per 100 PY in 2003 by 74.5% [71.6-77.1] to 0.36 [0.33-0.4] per 100 PY in 2018. However, among women aged 25-34, declines in HIV incidence were substantially slower (from 1.51 [1.45-1.57] per 100 PY in 2003 by 43.9% [40.5-47.4] to 0.84 [0.8-0.89] per 100 PY in 2018), and similarly in women aged 35-49 (from 0.9 [0.85-0.94] per 100 PY in 2003 by 37.4% [31.9-42.6] to 0.56 [0.52-0.6] per 100 PY in 2018), resulting in a progressive, substantial shift in the median age of infection in women from 23.4 [22.6-24.1] in 2003 to 28.2 [27.1-29.2] in 2018 (Fig. 1c-d). Progress in reducing HIV incidence thus continues to be substantially slower in women 14,21, especially among those aged 25 years and above. ![Fig. 2:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F2.medium.gif) [Fig. 2:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F2) Fig. 2: Time trends in age-specific sources of HIV infections in women and men. (**a**) Estimated age distributions of transmitting partners (posterior median: line, 95% credible interval: ribbon), along with the median age at transmission (posterior median: cross, 95% credible interval: linebar). Age contributions sum to 100% for each round, summing over men and women. (**b**) Estimated age distributions of transmitting partners by 5-year age bracket of infected partners (posterior median: thick black bar in boxplots, 50% interquartile range: height of box, 80% credible intervals: whiskers in boxplots). The width of the boxplots is proportional to the total infections in each recipient group. For reference, posterior estimates of the age distributions of sexual contact partners of men and women by 5-year age bands in the past 12 months in the same communities are shown in dark grey (estimates visualized in the same manner). (**c**) Comparison of the age contributions to transmitting partners (color) to the age contributions to men and women with unsuppressed HIV (posterior median: dashed black line, 95% credible interval: ribbon), along with median age (posterior median: cross, 95% credible interval: linebar). Age contributions sum to 100% for men and women combined. Throughout all subfigures, transmission flow estimates are based on *n* = 227 heterosexual source-recipient pairs identified among *n* =1,978 individuals in the transmission cohort and *n* =1,117 individuals in the incidence cohort. ### The proportion of transmission from men is increasing To characterize the population transmission flows by age and gender underlying observed shifts in incidence, we deep-sequenced virus from 1,978 participants with HIV (Supplementary Table S519). By embedding genomic surveillance into a population-based cohort study, deep-sequence sampling coverage was high relative to typical pathogen sequencing studies, which is essential for reconstructing transmission events 20,22,23,24,25. We characterized the phylogenetic ordering between multiple viral variants from individuals and estimated the direction of transmission with *phyloscanner* (Methods) 22,26. We identified 236 heterosexual source-recipient pairs that were phylogenetically close and exhibited, in combination with data on last negative and first positive tests, strongly consistent evidence of the direction of transmission (Methods and Extended Data Fig. 3). We further estimated the likely infection date from deep-sequence data 27, which enabled us to place the source-recipient pairs in calendar time and consider their age at the time of infection (Extended Data Fig. 4). Of the 236 heterosexual source-recipient pairs, we retained in total 227 pairs in whom transmission was estimated to have occurred during the study period. ![Fig. 3:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F3.medium.gif) [Fig. 3:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F3) Fig. 3: Changes in population-level suppression of HIV viral load. (**a**) Estimated trends in HIV prevalence and the proportion of census-eligible individuals in three age brackets that remain virally unsuppressed, defined as viral load above 1,000 copies/mL blood (posterior median: dots, 95% credible interval: errorbars), combining data from participants and from first-time participants as proxy of non-participants. (**b**) Male-to-female ratio in changes in population-level viral load suppression relative to round 10 (posterior median: dots, 95% credible interval: er-rorbars). (**c**) Estimated viral suppression rates by 1-year age band (x-axis) and gender (color) for survey round 18 (posterior median: dots, 95% credible interval: errorbars). Throughout all subfigures, estimates are based on data from *n* = 38, 749 participants including *n* = 3, 265 participants with HIV and with measured viral load. First-time participants were used as proxies of individuals who did not participate in the survey. ![Fig. 4:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F4.medium.gif) [Fig. 4:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F4) Fig. 4: Counterfactual modeling scenarios predicting the impact of interventions to increase HIV suppression in men on incidence reductions in women. (**a**-**b**) Estimated additional number of men with HIV in the census-eligible population in round 18 that already had suppressed virus (light grey), those who would have achieved viral suppression in the counterfactual intervention scenarios (color), and those who would have remained with unsuppressed virus in the counterfactuals (dark grey). Posterior median: bars, 95% credible interval: errorbars. (**c**) Percent reduction in incidence in women of the census-eligible population in round 18 under the counterfactual targeted scenarios. Posterior median: bars, 95% credible interval: errorbars. (**d**) Estimated incidence rates among women in the census-eligible population in round 18 (black solid line) and the counterfactual scenarios (color), with incidence rates among men in round 18 shown as reference (black dashed line). Posterior median: lines, 95% credible intervals: ribbons. Throughout all subfigures, estimates are based on data from *n* = 15, 053 participants in survey round 18, including *n* = 110 seroconverts in the incidence cohort in round 18, *n* = 432 individuals with HIV and with measured viral load in round 18, and *n* = 61 heterosexual source-recipient pairs in rounds 16-18, and information inferred through hierarchical models from all individuals in earlier rounds. Deep-sequence phylogenetics cannot prove direction of transmission between two persons 22, but in aggregate these data are able to capture heterosexual HIV transmission flows at a population level 20,28. We estimated population-level transmission flows adjusting for detection probabilities with semi-parametric Poisson flow regression models 29, and under the constraint that the transmission flows needed to closely match the age- and gender-specific incidence dynamics shown in Fig. 1 (Methods, Extended Data Fig. 5, and Supplementary Table S6). The fitted model was consistent with all the available data (Extended Data Fig. 6). Fig. 2a shows the age profile of the estimated male and female sources of infection, such that the male plus the female sources sum to 100% for each survey round. Overall, we found that the contribution of men to onward transmission increased progressively from 57.9% [56.2-59.6] in 2003 to 62.8% [60.2-65.2] in 2018, indicating that HIV transmission is now more disproportionately driven by men than has been the case previously. ### Transmissions from men are shifting to older ages The age profile of the population-level sources of infection characterizes the major age groups that sustain transmission 30. We find that the age of transmitting male partners progressively increased from a median age of 28.5 [27.1-30.1] years in 2003 to 33.5 [31.0-36.0] years in 2018 (Table 1 and Fig. 2a), and this increase in the age of transmitting male partners was largest in transmissions to women aged 20-24 (Fig. 2b). In contrast, the median age of female transmitting partners remained similar (from 25.0 [23.0-27.0] years in 2003 to 26.0 [24.0-28.0] years in 2018), corresponding to our earlier observations that the age of male incident infections also remained similar during the observation period. View this table: [Table 1:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T1) Table 1: HIV prevalence, viral suppression, transmission sources, and impact of counterfactual interventions focussed on closing the suppression gap in men by age of male partner, round 18, Oct 2016-Apr 2018. Over time, substantially fewer infections occurred in adolescent girls and young women aged 15-24 years. In 2003 the largest transmission flows were to women aged 15-24 years from male partners 0-6 years older (15.5% [12.3-18.9]) and from male partners 6+ years older (16.0% [12.7-19.2]) (Supplementary Table S7). By 2018, these transmission flows declined by approximately one third, with 8.1% [5.6-11.0], to women aged 15-24 years from male partners aged 0-6 years older, and 12.1% [9.3-15.2] to women aged 15-24 years from male partners aged 6+ years older. In those infections in adolescent girls and young women that occurred in 2018, the median age difference between incident infections in adolescent girls and young women and their transmitting male partners were 9.0 [7.0-12.0] years (Fig. 2b and Supplementary Table S7), similarly as in a phylogenetic study from KwaZulu-Natal in South Africa 31. This prompted us to estimate for comparison age-specific sexual contact patterns within RCCS communities (Methods and Supplementary Table S8). In 2018, the median age difference between adolescent girls and young women and their male sexual partners was 3.6 [3.5-3.9] years. Our data thus indicate that the main transmission flow into adolescent girls and young women is through contacts with considerably older men as compared to their typical sexual contacts 31,32, and that while this transmission flow has weakened overall, it remains the predominant mode of infection in adolescent girls and young women. By 2018, the largest share of transmission flows shifted to women aged 25-34 years, from male partners 0-6 years older. In 2003, transmissions to women 25-34 years from these transmitting partners accounted for 7.7% [6.2-9.3] of all transmissions, and by 2018 the share of these flows increased by half to 12.0% [9.1-15.0] (Supplementary Table S7). We also find that the transmission flows to women aged 35 years and above increased (Supplementary Table S7, also indicated by wider boxplots in Fig. 2b). Our data suggest further deviations in age-specific transmission flows from the typical sexual contact patterns within study communities. For all women aged 30 years and older, we estimate their male transmitting partners were of similar age with a posterior interquartile age range of 30.3-38.0 years in 2018, whereas for comparison the typical sexual contact partners of these women were older with a posterior interquartile age range of 40.0-42.7. These findings explain the unexpected age profile of male transmitting partners (Fig. 3c) that concentrates in men aged 25 to 40 instead of extending to progressively older men (Extended Data Fig. 7). Our observations are in line with recent studies from Zambia 20 and South Africa 33 that show having a male partner aged 25-40 years rather than the age gap between partners is associated with increased transmission risk. The transmission flows into men remained similar over time (Fig. 2b). In 2018, the largest transmission flow was to men aged 25-34 years from transmitting female partners of similar age that were 0-6 years older (10.6% [8.9-12.3]). ### Gender gaps in viral suppression are increasing We next placed the reconstructed shifts in transmission dynamics into the wider context of rapidly expanding HIV treatment during the observation period 14. We measured viral load from 2011 (survey round 15) among almost all participants with HIV (Supplementary Tables S1 and Extended Data Fig. 8) 34. Following WHO criteria 35, individuals with viral load measurements below 1,000 copies/millilitre (mL) plasma were considered viremic (Methods and Supplementary Table S9). By 2018, we find that the proportion of men and women who were viremic was entirely decoupled from HIV prevalence in that while the proportion of women with HIV was substantially higher than in men, the proportion of viremic women was similar or lower than in men (Fig. 3a). We quantified these trends with the male-to-female ratio of the proportion of viremic individuals relative to 2003 levels, which has been progressively increasing in all age groups (Fig. 3b). This suggests 36 that faster rises in female HIV suppression could explain in part the faster declines in male incidence rates as higher rates of ART uptake and virus suppression in women mean that male partners are less likely to become infected, whereas men’s higher rates of unsuppressed virus mean they are more likely to transmit to female partners (Extended Data Fig. 9). These trends have by 2018 accumulated to a substantial gap in suppression levels in men compared to women (Table 1 and Fig. 3c). ### Men contribute disproportionally to transmission Combining phylogenetics with the virus suppression data also allowed us to compare transmission with population-level infectiousness as measured through viremic individuals (Table 1 and Fig. 2c). In 2018, the contribution of men to viremic individuals was (49.2% [44.3-54.1]). For the same time period we found that the contribution of men to transmission was consistently higher (62.8% [60.2-65.2]), indicating that men contribute more to transmsission than population-viral load suggests. These findings are compatible with generally higher viral load in men than women 34,37 that are expected to lead to higher transmission rates per sex act from men than women, heterogeneous contact patterns 38, higher biological susceptibility of women to HIV infection in heterosexual contacts 39,40, but also lower susceptibility of men to HIV infection following voluntary medical male circumcision 41. ### Policy implications It has been previously demonstrated that people with HIV who are on ART and maintain suppressed virus do not transmit HIV 42,43. On this basis, we quantified the impact that closing the gap in male-female virus suppression levels could have had on HIV transmission flows. Specifically, we parameterised the transmission flow model in terms of HIV seronegative individuals who are susceptible to infection and individuals with unsuppressed HIV who remain infectious. Thus, we were able to use the fitted model to estimate the impact of fewer individuals with unsuppressed HIV on evolving HIV transmission in counterfactual, modelled intervention scenarios (see Methods). We considered the impact of three hypothetical scenarios: first, the impact of reducing by half the gap in the proportion of men with suppressed virus as compared to women (“closing half the suppression gap in men”) at the end of the observation period in 2018 (Fig. 3c); second, the impact of achieving the same virus suppression levels in men with HIV as in women in 2018 (“closing the suppression gap in men”); and third—for reference—achieving the UNAIDS 95-95-95 target that 86% of men (0.95 * 0.95 * 0.95) with HIV reach viral suppression in all age groups in 201844. Table 1 and Fig. 4a describe the age-specific male counterfactual viral suppression targets of each scenario, and place these into the context of prevalence, suppression, and transmission. Overall, we found slightly older men would have reached suppression in the scenarios closing the suppression gap as compared to the UNAIDS 95-95-95 scenario. We predict that in the UNAIDS 95-95-95 scenario, an additional 172.6 [136.8-210.0] men with HIV would have reached viral suppression in 2018 (Fig. 4b) and this would have resulted in a 58.4% [54.9-61.7] additional reduction in HIV incidence in women in 2018 (Fig. 4c), which is in good agreement with the contribution of 95-95-95 interventions to projected incidence reductions for all of Eastern and Southern Africa under the mathematical models used to inform the global HIV prevention strategy 45. In the scenario closing half the suppression gap in men, an additional 75.1 [53.9-96.0] men with HIV would have reached viral suppression in 2018 and resulted in a 25.1% [24.2-26.2] additional reduction in HIV incidence in women in 2018. In the scenario closing the entire suppression gap in men, an additional 150.2 [107.8-193.0] men with HIV would have reached viral suppression in 2018 and resulted in a 50.6% [48.6-52.8] additional reduction in HIV incidence in women in 2018 (Fig. 4b-c). Thus, all three intervention scenarios involved reaching a small additional number of men compared to the thousands of women with higher risk of HIV acquisition in the same rural and semi-urban study areas 46. We predict that closing the suppression gap in men would have changed the female-to-male incidence rate ratio from 1.59 [1.38-1.82] to 0.78 [0.69-0.87] in 2018 (Fig. 4d), entirely closing the growing gender disparity in HIV incidence. ## Discussion Effective HIV interventions and services are essential to bring most African countries on track to end AIDS as a public health threat by 2030 and accelerate progress to-wards the vision of the UNAIDS “three Zeros” target: zero new HIV infections, zero discrimination, and zero AIDS-related deaths 45,47. Gender inequalities are among the main reasons why global targets on mass scale-up of HIV testing, biomedical interventions and on incidence reductions have not been achieved 48. Here, we combined population-based incidence with deep-sequence viral phylogenetic surveillance data to characterize how HIV incidence and heterosexual transmission sources have been changing by age and gender in a typical rural and semi-urban African setting. We show that along with increasing availability of HIV services, there have been consistently faster increases in viral suppression in women than men and an increasing majority of new infections are arising from men. We also document substantial age shifts in HIV incidence and transmission sources, with the primary burden of incidence shifting to older women aged 25-34 years, the primary burden of transmission shifting to male partners aged 30-39 years, and the relative contribution of transmission flows to adolescent girls and young women from older men reducing by one third. Modeling counterfactual improvements in HIV outcomes for men on the inferred transmission flows during the last survey round in 2016-2018, we find that closing the male gender gap in viral suppression rates could have reduced incident female infections by half in that time period and brought about gender equality in HIV infection burden. This study evaluated data from one longitudinal surveillance cohort in southern Uganda, but the increasing gender disparities and shifts in age-specific transmission are not unique. Incidence data published over the last decade documents widespread declining incidence across the African continent 17, greater differences in rates of new infections between men and women over calendar time, and rising average age of infection in women 17. Data from population surveillance studies and HIV treatment and prevention trials shows higher levels of viremia among men compared to women with HIV 49,50, and phylogenetic studies from Botswana 51 and Zambia 20 also report gender disparities in HIV transmission. Together, these observations suggest that the principal characteristics of the evolving HIV epidemic likely apply more broadly in similar rural and semi-urban populations across Eastern and Southern Africa. Given that the African HIV epidemic has historically been concentrated among adolescent girls and young women 4,5, programs and policies rightfully have concentrated on reducing HIV risk in this demographic. Here, we document that most heterosexual transmission is driven by men and that — as incidence is declining — the contribution of men to onward heterosexual transmission is growing, likely due to slower population-level declines in HIV viremia in men. While there are emerging efforts to design male-centered HIV interventions 52,53, African men continue to be overlooked in the design of programmatic services 54,55. Many factors, including gender norms, mobility, and lack of targeted programming to men contribute to lower uptake of HIV services by men 53. Case finding of men with HIV might be difficult but could be strengthened by expanding access to HIV testing services most likely to reach them, such as through self-testing or assisted partner notification and other social network strategies 54,56,57. Retention of men with HIV in treatment and care programs could be improved through male-centered differentiated service delivery. It is well-established that improving male engagement in HIV services leads to better health for men 58,59. We expect additional interventions such as voluntary medical male circumcision, condom promotion, or pre-exposure prophylaxis would lead to further reductions in new cases 60. Our findings are grounded in fifteen years of consecutive population-based epidemiologic and molecular surveillance in southern Uganda, enabling us to measure changes in HIV incidence and transmission during a critical period of HIV service scale-up. Though it is typically assumed that age-specific patterns in onward HIV transmission correspond to those of viremia or follow typical sexual contact patterns, we find that this is not always the case. First, men contributed disproportionally more to onward heterosexual transmission than to viremia across all survey rounds during which viremia were measured (Fig. 2c and Extended Data Fig. 7a). Second, older women contributed less to transmission than viremia suggest, an observation that was consistent with attenuating sexual activity of women from age 25 onwards (Extended Data Fig. 7a). Third, young women and young men tended to be infected by transmitting partners who were substantially older than the typical sexual partners of the same population age group (Fig. 2b and Extended Data Fig. 7b). These findings illustrate the central utility of pathogen genomics to track and understand patterns of transmission, especially when interpreted in the context of population-based surveillance data, and when implemented at high enough sequence coverage to reconstruct directed transmission networks. This study has important limitations. First, not all census-eligible individuals participated in the survey, primarily due to absence for work or school (Extended Data Fig. 1) 14. We used data from first-time participants as proxies of non-participants, but we cannot rule out that non-participants include disproportionally larger populations of people with HIV and/or with different risk profiles. In this case, sensitivity analyses (Supplementary Table S10) indicate that more viremic men would have to be reached in all intervention scenarios for similar HIV incidence reductions in women as in Fig. 4. Second, we were only able to deep-sequence a fraction of all transmission events, and these may not be representative of all transmissions. We characterized sampling probabilities under the assumption that individuals were ever deep-sequenced at random within age- and gender strata, and found that the sampling probabilities did not differ substantially between strata in each round (Extended Data Fig. 5), so that the estimated transmission flows were not sensitive to our sampling probability adjustments (Supplementary Table S10). Of course, these sampling adjustments are modeled and it remains possible that missing data could bias our findings. Third, our error analyses indicate that deep-sequence phylogenetics are not a perfect marker of direction of transmission, with estimated false discovery rates of 16.3% [8.8-28.3%] in this cohort. 22 Fourth, over time some communities were added and others left the Rakai Community Cohort Study (Supplementary Table S2). We repeated our analysis on the subset of 28 continuously surveyed communities, and found similar incidence and transmission dynamics (Supplementary Table S10). Fifth, our findings on rural and semi-urban populations may not extend to populations with different demographics, risk profiles or healthcare access, and this includes populations in urban or metropolitan areas or key populations. This study demonstrates shifting patterns in HIV incidence and in the drivers of HIV infection in communities typical of rural and semi-urban East Africa, providing key data for evidence-informed policy making. We find incidence rates have dropped substantially in women aged 15-24 years from 2003 to 2018, and incidence rates now peak among women aged 25-35 years, consistent with cross-sectional national surveillance data from Uganda 61. Shifts in women’s incidence are the result of an increase in the age of transmitting male partners, and the primary contribution to HIV transmission lies now in men aged 30 and above. The growing contribution of men to heterosexual transmission is associated with substantially slower declines of unsuppressed viremia in men than women. We predict successful interventions centered on men that bring suppression rates in men on par with those in women could reduce incidence in women by half and close the gender gap in new infections. These findings reinforce calls for HIV prevention programming and services to give greater priority to reach and retain in care men with HIV as this will improve male health, substantially reduce incidence in women, and close gender gaps in infection burden. ## Methods ### Rakai Community Cohort Study #### Longitudinal surveillance Between September 2003 and May 2018, nine consecutive survey rounds of the Rakai Community Cohort Study (RCCS) were conducted in 36 inland communities in south-central Uganda (Fig. 1, Supplementary Tables S1-S2, and Supplementary Fig. S1). The results presented in this paper derive from data collected through these surveys, including the population census, the RCCS survey participants, the incidence cohort, and the phylogenetic transmission cohort. RCCS survey methods have been reported previously 14,18. In brief, for each survey round, the RCCS did a household census, and subsequently invited all individuals that were of age 15-49 years and residents for at least 1 month to participate in the open, longitudinal RCCS survey; and so data collection was not randomized, and data collection was blind relative to previous interactions with individuals or any personal characteristics apart from age and residency status, and any research questions. Eligible individuals first attended group consent procedures, and individual consent was obtained privately by a trained RCCS interviewer. Following consent, participants reported in a private location, typically a tent at the survey hub, on demographics, behavior, health, and health service use. All participants were offered free voluntary counseling and HIV testing as part of the survey. Rapid tests at the time of the survey and confirmatory enzyme immunoassays were performed to determine HIV status. All participants were provided with pre-test and post-test counseling, and referrals of individuals who were HIV-positive for ART. Additionally, all consenting participants, irrespective of HIV status, were offered a venous blood sample for storage/future testing, including viral phylogenetic studies. Supplementary Table S1 summarises the characteristics of the RCCS participants and HIV-positive participants by age and gender. For the purpose of our analyses, we combined data from three pairs of geographically close areas in peri-urban settings into three communities, and 28 of 36 communities were continuously surveyed over all rounds (Supplementary Table S2). All epidemiologic data collected through RCCS are stored in a database running Microsoft SQL server 2019 and Microsoft Access version 2016. #### Ethics declarations The study was independently reviewed and approved by the Ugandan Virus Research Institute, Scientific Research and Ethics Committee, protocol GC/127/13/01/16; the Ugandan National Council of Science and Technology; and the Western Institutional Review Board, protocol 200313317. All study participants provided written informed consent at baseline and follow-up visits using institutional review board approved forms. This project was reviewed in accordance with CDC human research protection procedures and was determined to be research, but CDC investigators did not interact with human subjects or have access to identifiable data or specimens for research purposes. Participants in the RCCS received 10,000UGX (approximately 2.50USD) in compensation for the baseline and follow-up surveys. #### Population size estimates To characterize changes in population demography, individual-level data on the census-eligible individuals that were obtained during each census were aggregated by gender, 1-year age band (between 15 and 49 years) and survey round (Extended Data Fig. 1a-b, bars). The age reported by household heads in the census surveys tended to reflect grouping patterns towards multiples of 5, suggesting that household heads reported ages only approximately. For this reason, we smoothed population sizes across ages independently for every gender and survey round, using locally weighted running line smoother (LOESS) regression methods that fit multiple polynomial regressions in local neighborhoods as implemented in the R package stats version 3.6.2 with span argument set to 0.5 (Extended Data Fig. 1a-b, line). Model fit was assessed visually without a formal test, suggesting that the data met the assumptions of the statistical model. #### Participation rates To characterize participation rates, we calculated the proportion of RCCS participants in the census-eligible population by gender, 1-year age band and survey round (Extended Data Fig. 1c-d, bars). Following consent, participants reported either their birth date or current age themselves, and accompanying documentary evidence was requested. There were no obvious age grouping patterns of multiple of 5 among participants. Overall, participation rates were lower in men than women (63% vs. 75%). Participation rates also increased with age for both men and women, and were very similar across survey rounds. Considering the grouping patterns by age in the population count data, we again smoothed the participation rates across ages independently for every gender and survey round using LOESS regression as specified above for population size estimation (Extended Data Fig. 1c-d, line). Model fit was assessed visually without a formal test, suggesting that the data met the assumptions of the statistical model. #### HIV status and prevalence All RCCS participants were offered free HIV testing. Prior to October 2011, HIV testing was performed through enzyme immunoassays (EIAs) with confirmation via Western Blot and DNA PCR. After October 2011, testing was performed through a combination of three rapid tests with confirmation of positives, weakly positives and discordant results by at least two EIAs and Western Blot or DNA PCR 62. Overall, 99.7% participants took up the test offer across survey rounds, and Supplementary Table S1 documents the number of participants with HIV. From these survey data, we estimated HIV prevalence (i.e., probability for a participant to have HIV) with a non-parametric Bayesian model over the age of participants independently for both genders and survey round. Specifically, we used a binomial likelihood on the number of participants with HIV parameterized by the number of participants and HIV prevalence in each 1-year age band. The HIV prevalence parameter was modeled on the logit scale by the sum of a baseline term and a zero-mean Gaussian Process on the age space. The prior on the baseline was set to a zero-mean normal distribution with a standard deviation of 10. The covariance matrix of the Gaussian Process was defined with a squared exponential kernel, using a zero-mean half-normal prior with a standard deviation of 2 on the scale parameter of the squared exponential kernel and a zero-mean half-normal prior with a standard deviation of 11.3 ((49 − 15)*/*3) on the lengthscale of the squared exponential kernel. The model was fitted with Rstan release 2.21.0 using Stan’s adaptive Hamiltonian Monte Carlo (HMC) sampler 63 with 10,000 iterations, including warm-up 500 iterations. Convergence and mixing were good, with highest R-hat value of 1.0029, and lowest effective sample size of 830). The model represented the data well, with 98.57% of data points inside 95% posterior predictive intervals, indicating that the data met the assumptions of the statistical model. For the mathematical modeling of transmission flows, we assumed that age- and gender-specific HIV prevalence were the same in non-participants in the RCCS communities as in the participants in these communities. #### ART use The RCCS measures ART use through participant reports since survey round 11. Self-reported ART use reflected viral suppression with high specificity and a sensitivity around 70% in the study population (Supplementary Table S9). We took the following pre-processing steps. For survey round 10, we assumed self-reported ART use to have been on the same levels as in round 11. Next, the ART use field was adjusted to “yes” for the participants with HIV who did not report ART use but who had a viral load measurement below 1,000 copies per milliliter (mL) plasma blood. Further, we considered it likely that with increasingly comprehensive care and changing treatment guidelines 14,64, ART use in individuals with HIV who did not participate increased substantively over time, and this prompted us to consider as proxy of ART use in non-participants the observed ART use in first-time participants with HIV. Overall, first-time participants represented between 15.26% to 39.87% of all participants. Extended Data Fig. 8a-b exemplifies the self-reported ART use data in male participants and male first-time participants, along with the combined estimate of individuals with HIV in the study population who report ART use, summing over participants and non-participants. These estimates were obtained using the same Bayesian non-parametric model as for HIV prevalence. Convergence and mixing were good, with highest R-hat value of 1.0025 and lowest effective sample size of 978 for the participants and 1.0027, 521 respectively for first-time participants. The model represented the data well, with 99.67% of data points for the participants inside the corresponding 95% posterior predictive intervals, and 99.24% for the first-time participants, indicating that the data met the assumptions of the statistical model. The resulting, estimated ART use rates in infected men and women are shown in Extended Data Fig. 8c. #### Viral suppression Since survey round 15, HIV-1 viral load was measured on stored serum/plasma specimens from infected participants using the Abbott real-time m2000 assay (Abbott Laboratories, IL, USA), which is able to detect a minimum of 40 copies/mL. Viral suppression was defined as a viral load measurement below 1,000 copies/mL plasma blood following recommendations of the World Health Organisation (WHO) 35. To estimate virus suppression levels in the infected non-participants, we considered again as proxy data on infected first-time participants. Overall, viral load measurements were obtained from 19.3% of participants with HIV in survey round 15 and nearly all (>97.71%) participants with HIV since survey round 1665,66,67. From these data we estimated the proportion of individuals in the study population with HIV who had suppressed virus, summing over participants and non-participants, using the same Bayesian non-parametric model as for HIV prevalence and ART use. Convergence and mixing were good with lowest R-hat value of 1.0016 and lowest effective sample size of 461 for the participants and 1.0052, 844 respectively for the first-time participants. The model represented the data well, with 98.19% of data points inside 95% posterior predictive intervals and 97.99% for the first-time participants, indicating that the data met the assumptions of the statistical model. For the purpose of mathematical modeling of transmission flows, we next considered the earlier survey rounds 10 to 14, for which viral load measurements were not available. On average, 93% of individuals reporting ART use also had suppressed virus (Supplementary Table S9), leading us to estimate the number of individuals with suppressed virus before 2011 from corresponding ART use data. Specifically, we estimated the proportion of the study population with HIV that was virally suppressed by adjusting the estimated ART use data with the sensitivity of being virally suppressed given self-reported ART use and the specificity of being virally suppressed given self-reported no ART use estimated from round 15 where available, and otherwise from round 16 (Supplementary Table S9). Specificity and sensitivity values by 1-year age bands were linearly interpolated between the midpoints of the age brackets in Supplementary Table S9. The resulting, estimated virus suppression levels in men and women with HIV are shown in Extended Data Fig. 8d, illustrating that the gap in virus suppression levels increased over time. #### Sexual behavior RCCS participants reported to interviewers in each round on aspects of sexual behavior, including the number of sexual partners in the past 12 months within the same community, the number of partners outside the community, and in round 15 also demographic characteristics of up to four partners (Supplementary Table S8). To interpret HIV transmission flows in the context of typical sexual contact networks, we focused on the detailed behavior data collected in round 15 and estimated sexual contact intensities between men and women by 1-year age band, defined as the expected number of sexual contacts of one individual of gender *g* and age *a* with the population of the opposite gender *h* and age *b* in the same community. Estimates were obtained with the Bayesian rate consistency model, version 1.0.0, using default prior specifications 68. We noted along with previous work 69,70,71,72 that women tended to report considerably fewer contacts than men (Supplementary Table S8), prompting us to include in the linear predictor of contact rates additional age-specific random effects to capture under-reporting behavior in women. Further, community-specific baseline parameters were added to allow for variation in the average level of contact rates in each community, but the age-specific structure of contact rates was assumed to be identical across communities. The resulting model was fitted to all data pertaining to within-community sexual contacts in the last year, including reports of within-community contacts for which information on the partners remained unreported. Contacts reported with partners from outside the same community were excluded, because male-female contacts have to add up to femalemale contacts only in the same population denominator, and hence under-reporting could only be adjusted for when within-community contacts are considered. The model was fitted with CmdstanR version 0.5.173 using Stan’s adaptive HMC sampler 63 with 4 chains, where each chain runs 2800 iterations, including 300 warm-up iterations. Convergence and mixing were good, with highest R-hat value of 1.003, and lowest effective sample size of 1,745. The model represented the data well, with *>* 99% of data points inside 95% posterior predictive intervals, indicating that the data met the assumptions of the statistical model. Supplementary Table S8 reports the estimated sexual contact intensities from men and women in survey round 15, and shows that the estimated, under-reporting adjusted sexual contact intensities in women were considerably higher than those directly reported. The table also shows that the estimated number of sexual contacts from men to women equal those from women to men, and the estimated age distribution of sexual contacts is shown in Fig. 2 and Extended Data Fig. 7. ### Longitudinal HIV incidence cohort #### Data and outcomes from the incidence cohort The RCCS encompasses both a full census of the study communities and a population-based survey in each surveillance round, which enables identification and follow up of unique individuals over time, and thus provides a comprehensive sampling frame to measure HIV incidence. The RCCS incidence cohort comprises of all RCCS study participants who were HIV-negative at their first visit (baseline) and had at least one subsequent follow-up visit (Supplementary Fig. S1). Individuals in the incidence cohort were considered to be at risk of acquiring HIV after their first visit, and stopped accruing risk at the date of HIV acquisition or the date of last visit. Exposure times were estimated from data collected at survey visit times similarly as in 14. Individuals in the incidence cohort who remained negative until the last survey round contributed their time between the first and last survey visit to their exposure period. Individuals in the incidence cohort who were found to have acquired HIV must have done so between the visit date of the last round in which they were negative and the visit date of the current round, and the infection date was imputed at random between the two dates. This included incident cases who had no missed visit between the last negative and current visit (type 1) or one missed visit (type 2) as in 14, but also cases who had more than one missed visit (type 3). Unknown dates were imputed at random 50 times, and individual exposure periods and incident cases were then attributed to each survey round, summed over the cohort, and then averaged over imputations. Supplementary Table S3 and Extended Data Fig. 2 illustrate the age- and gender-specific exposure times and incidence events in each survey round. In sensitivity analyses, we considered only those individuals in the incidence cohort who resided in one of the 28 inland communities that were continuously surveyed across all rounds 10 to 18, and found similar incidence dynamics with slightly faster declines in incidence rates in younger men, although this difference was not statistically significant. No statistical methods were used to pre-determine sample sizes but our sample sizes are similar to those reported in previous publications 14. #### Modelling and analysis The primary statistical objective was to estimate longitudinal age-specific HIV incidence rates by 1-year age bands across (discrete) survey rounds, separately for each gender. We used a log-link mixed-effects Poisson regression model, with individual-level exposure times specified as offset on the log scale, common baseline fixed effect, and further random effects. The random effects comprised a one-dimensional smooth function on the age space, a one-dimensional smooth function on the survey round space, and an interaction term between age and survey round. The functions were specified as one-dimensional Gaussian processes, similar as in the model for estimating HIV prevalence. Alternative specifications, including two-dimensional functions over the participant’s age and survey round, and without interaction terms between age and survey rounds were also tried. We did not consider incidence trends in continuous calendar time because study communities were surveyed in turn, and so the incidence data within each round are structured by communities, which would require further modeling assumptions to account for. Due to the large number of individual observations, models were fitted using maximum-likelihood estimation (MLE) with the R package mgcv version 1.8-38 in the R language 74, to each of the 50 data sets with imputed exposure times for each gender. Numerical convergence was examined with the gam.check function. Within and between sample uncertainties in parameter estimates, from the variability of the estimation procedure and the data imputation procedure, were incorporated in the age-, gender- and survey round-specific incidence rate estimates by drawing 1,000 replicate incidence rate estimates from the MLE model parameters and associated standard deviation obtained on each of the 50 imputation data sets, and then calculating median estimates and 95% prediction intervals over the 1, 000 × 50 Monte Carlo estimates (Fig. 1c). Model fits were evaluated by comparing predicted HIV incidence infections estimates to the empirical data. To assess model fit, incident cases were predicted using the Poisson model parameterised by replicate MLE incidence estimates. Overall, model fit was very good, with 98.80% [98.10-99.49] data points inside the 95% prediction intervals across all 50 imputed data sets and the fitted model was consistent with all the available data (Extended Data Fig. 6), indicating that the data met the assumptions of the statistical model. The Akaike information criterion was used to identify the best model for each gender, and the best model was as described above (Supplementary Table S4). ### Longitudinal viral phylogenetic transmission cohort #### Data from the transmission cohort Within the RCCS, we also performed population-based HIV deep-sequencing spanning a period of more than 6 years, from January 2010 to April 2018. The primary purpose of viral deep sequencing was to reconstruct transmission networks and identify the population-level sources of infections, thus complementing the data collected through the incidence cohort. The RCCS viral phylogenetic transmission cohort comprises of all participants with HIV for whom at least one HIV deep sequence sample satisfying minimum quality criteria for deep-sequence phylogenetic analysis is available (Supplementary Fig. S1). For survey rounds 14 to 16 (PANGEA-HIV 1), viral sequencing was performed on plasma samples from participants with HIV who had no viral load measurement and self-reported being ART-naïve at the time of the survey, or who had a viral load measurement above 1,000 copies/mL plasma. We used this criterion because viral deep sequencing was not possible within our protocol on samples with virus less than 1,000 copies/mL plasma, and because self-reported ART use was in this population found to be a proxy of virus suppression with reasonable specificity and sensitivity 14,22. Plasma samples were shipped to University College London Hospital, London, United Kingdom, for automated RNA sample extraction on QIAsymphony SP workstations with the QIAsymphony DSP Virus/ Pathogen Kit (Cat. No. 937036, 937055; Qiagen, Hilden, Germany), followed by one-step reverse transcription polymerase chain reaction (RT-PCR) 75. Amplification was assessed through gel electrophoresis on a fraction of samples, and samples were shipped to the Wellcome Trust Sanger Institute, Hinxton, United Kingdom for HIV deep-sequencing on Illumina MiSeq and HiSeq platforms in the DNA pipelines core facility. Primers are publicly available 75. For survey rounds 17 to 18 (PANGEA-HIV 2), viral load measurements were available for all infected participants and viral sequencing was performed on plasma samples of individuals who had not yet been sequenced and who had a viral load measurement above 1,000 copies/mL plasma. Plasma samples were shipped to the Oxford Genomics Centre, Oxford, United Kingdom, for automated RNA sample extraction on QIAsymphony SP workstations with the QIAsymphony DSP Virus/ Pathogen Kit (Cat. No. 937036, 937055; Qiagen, Hilden, Germany), followed by library preparation with the SMARTer Stranded Total RNA-Seq kit v2 - Pico Input Mammalian (Clontech, TaKaRa Bio), size selection on the captured pool to eliminate fragments shorter than 400 nucleotides (nt) with streptavidin-conjugated beads 76 to enrich the library with fragments desirable for deep-sequence phylogenetic analysis, PCR amplification of the captured fragments, and purification with Agencourt AMPure XP (Beckman Coulter), as described in the veSEQ-HIV protocol 77. Sequencing was performed on the Illumina NovaSeq 6000 platform at the Oxford Genomics Centre, generating 350 to 600 base pair (bp) paired-end reads. Sequencing probes are publicy available 78. A subset of samples from survey rounds 14 to 16 with low quality read output under the PANGEA-HIV 1 procedure was re-sequenced with the veSEQ-HIV protocol. To enhance the genetic background used in our analyses, additional samples from the spatially neighboring MRC/UVRI/LSHTM surveillance cohorts and other RCCS communities were also included. For sequencing, the following software were used, QuantStudio Real-Time PCR System v1.3, Agilent TapeStation Software Analysis 4.1.1, Clarity Version 4.2.23.287, FreezerPro 7.4.0-r14598, and LabArchives Electronic Lab Notebook 2023. We restricted our analysis to samples from 2,172 individuals that satisfied minimum criteria on read length and depth for phylogeny reconstruction and sub-sequent inferences. Specifically, deep-sequencing reads were assembled with the shiver sequence assembly software, version 1.5.779. Next, phyloscanner version 1.8.126 was used to merge paired-end reads, and only merged reads of at least 250 bp in length were retained in order to generate 250bp deep-sequence alignments as established in earlier work 22. Deep-sequencing was performed from 2010 (survey round 14) onwards, but because sequences provide information on past and present transmission events, we also obtained information on transmission in earlier rounds and calculated sequence coverage in participants that were ever deep-sequenced at minimum quality criteria for phylogenetic analysis. Specifically, we required that individuals had a depth of ≥30 reads over at least 3 non-overlapping 250bp genomic windows. Individuals who did not have sequencing output meeting these criteria were excluded from further analysis, and these were largely individuals sequenced only in PANGEA-HIV 1, and were primarily associated with low viral load samples 77,80. In total, we deep-sequenced virus from 1,978 participants with HIV of who 559 were also in the incidence cohort. Supplementary Table S5 characterizes HIV deep-sequencing outcomes in more detail. No statistical methods were used to pre-determine sample sizes but our sample sizes are similar to those reported in previous publications 20,28,51. #### Reconstruction of transmission networks and source-recipient pairs The HIV deep-sequencing pipeline provided sequence fragments that capture viral diversity within individuals, which enables phylogenetic inference into the direction of transmission from sequence data alone 22,79,81. First, potential transmission networks were identified, and in the second step transmission networks were confirmed and the transmission directions in the networks were characterized as possible. In this study, the first step was modified from previous protocols 22 to ease computational burden, while the second step was as before performed with phyloscanner, using version 1.8.1. In the first step 82, to identify potential transmission networks, HIV consensus sequences were generated as the most common nucleotide in the aligned deep-sequence fragments that were derived for each sample. We then calculated similarity scores between all possible combinations of consensus sequences in consecutive 500 bp genomic windows rather than the entire genome to account for the possibility of recombination events and divergent virus in parts of the genome. Similarity score thresholds to identify putative, genetically close pairs were derived from data of long-term sexual partners enrolled in the RCCS cohort similarly as in 22,82, and then applied to the population-based sample of all possible combinations of successfully sequenced individuals. Overall, 2525 putative, genetically close individuals were identified, and these formed 305 potential transmission networks. In the second step, we confirmed the potential transmission networks in phylogenetic deep-sequence analyses. We updated the background sequence alignment used in phyloscanner to a new sequence data set that included 113 representatives of all HIV subtypes and circulating recombinant forms and 200 near full-genome sequences from Kenya, Uganda, and Tanzania, obtained from the Los Alamos National Laboratory HIV Sequence Database ([http://www.hiv.lanl.gov/](http://www.hiv.lanl.gov/)). The deep-sequence alignment options were updated to using MAFFT version 7.475 with iterative refinement 83, and additional iterative re-alignment using consistency scores in case a large proportion of gap-like columns in the first alignment was detected. Deep-sequence phylogeny reconstruction was updated to using IQ-TREE version 2.0.3 with GTR+F+R6 substitution model, resolving the previously documented deep-sequence phylogenetics branch length artefact 20,84. Confirmatory analyses of the potential transmission networks were updated to using phyloscanner version 1.8.1 with input argument *zeroLengthAdjustment* set to TRUE. From phyloscanner output, we calculated pairwise linkage scores that summarise how frequently viral phylogenetic subgraphs of two individuals were adjacent and phylogenetically close in the deep-sequence phylogenies corresponding to all 250bp genomic windows that contained viral variants from both individuals 22,26. Similarly we calculated pair-wise direction scores that summarise how frequently viral phylogenetic subgraphs of one individual were ancestral to the subgraphs of the other individual in the deep-sequence phylogenies corresponding to all 250bp genomic windows that contained viral variants from both individuals and in which subgraphs had either ancestral or descendant relationships 22,26. Phylogenetically likely source-recipient pairs with linkage scores ≥ 0.5 and direction scores ≥ 0.5 were extracted, and only the most likely source-recipient pair with highest linkage score was retained if multiple likely sources were identified for a particular recipient. The resulting source-recipient pairs were checked further against sero-history data from both individuals where available. If sero-history data indicated the opposite direction of transmission, the estimated likely direction of transmission was set to that indicated by sero-history data. #### Infection time estimates The shape and depth of an individual’s subgraph in deep-sequence phylogenies also provide information on the time since infection, and since the sequence sampling date is known thus also on the infection time 85 and the age of both individuals at the time of the infection event. We used the phyloTSI random forest estimation routine with default options, which was trained on HIV serocon-verter data from the RCCS and other cohorts, and uses as input the output of the phyloscanner software 27. Individual-level time since infection estimates were associated with wide uncertainty (Extended Data Fig. 4a), and for this reason we refined estimates for the phylogenetically likely recipient in source-recipient pairs using the inferred transmission direction, age data, and where available longitudinal sero-history data. Specifically, we refined plausible infection ranges as indicated in the schema in Supplementary Fig. S2. Here, the dotted red rectangle illustrates the 2.5% and 97.5% quantiles of the phyloTSI infection time estimates for the phylogenetically likely recipient (x-axis) and transmitting partner (y-axis). We incorporated evidence on the direction of transmission by requiring that the date of infection of the phylogenetically likely recipient is after that of the transmitting partner (filled red triangle). Sero-history and demographic data were incorporated as follows. For both the recipient and the transmitting partner, the upper bound of the infection date was set as the 30th day prior to the first positive test of the participant 86. The lower bound of the infection date was set to the largest of the following dates, the date of last negative test if available, the 15th birthday, or the date corresponding to 15 years prior the upper bound 87. The refined uncertainty range of the infection time estimates of the phylogenetically likely transmitting partner and recipient are illustrated as the purple triangle in the schema above, and obtained as follows. Firstly, we defined individual-level plausible ranges, by intersecting the range of dates consistent with the phyloTSI predictions and sero-history data. If the intersection was empty, we discarded the phyloTSI estimates. Then we intersected the rectangle given by the cartesian product of the plausible intervals for source and recipient with the half-plane consistent with the direction of transmission. Finally, infection dates were sampled at random from the refined uncertainty range, so that the median infection date estimates correspond to the center of gravity of the triangle (cross). In sensitivity analyses, we further integrated estimates of transmission risk by stage of infection 88, though this had limited impact on the estimates (see Sensitivity analyses section below). In cases where the likely transmitting partner in one heterosexual pair was the recipient partner in another heterosexual pair, the above infection date refinement algorithm was applied recursively so that the refined infection date estimates were consistent across pairs. Finally, the transmission events captured by each source-recipient pair were attributed to the survey round into which the posterior median infection time estimate of the recipient fell, and in cases where the median estimate fell after the start time of a round and the end time of the preceding round, the event was attributed to the preceding round. In total, we identified 539 source-recipient pairs that involved participants from the 36 survey communities and further individuals from the background data set. In 13 of the 539 source-recipient pairs, available dates of last negative tests indicated that only the opposite transmission direction was possible and in these cases the inferred direction of transmission was set to the opposite direction. The resulting pairs included 501 unique recipient partners, and for reach we retained the most likely transmitting partner. To identify pairs capturing transmission events within the RCCS inland communities, we restricted analysis initially to 236 heterosexual source-recipient pairs in whom both individuals were ever resident in the 36 survey communities. Of these, 142 pairs were from men to women and 94 from women to men. Infection times were estimated for all sampled individuals and refined for the recipient partners in the 236 heterosexual source-recipient pairs. For 4 recipient partners, the phyloTSI estimates were ignored as they were incompatible with inferred transmission direction and survey data, and was based on sero-history data only. The phylogenetically most likely location of both individuals at time of transmission was estimated as their location at the RCCS visit date that was closest to the posterior median infection time estimate. Using this location estimate, 233 of the 236 heterosexual source-recipient pairs were estimated to capture transmission events in RCCS inland communities and were retained for further analysis. A further 6 recipient partners had posterior median infection time estimates outside the observation period from September 2003 to May 2018 and were excluded, leaving for analysis 227 heterosexual source-recipient pairs that captured transmission events in RCCS inland communities during the observation period. This excluded 88 potential source-recipient pairs from our study due to ethical considerations and prior analyses suggesting these pairs most likely represent partially sampled transmission chains (i.e., “false positives”) 22. ### Transmission flow analysis #### Statistical framework We next estimated the sources of the inferred population-level HIV incidence dynamics from the dated, source-recipient pairs in the viral phylogenetic transmission cohort. Overall, inference was done in a Bayesian frame-work using a semi-parametric Poisson flow model similar to Xi, X. *et al*. 29, that was fitted to observed counts of transmission flows ![Graphic][1] with transmission direction *g* ⟶ *h* (male-to-female or female-to-male), time period *p* (R10-R15 and R16-R18) in which the recipient was likely infected, and 1-year age bands *i, j* of the source and recipient populations respectively, where ![Formula][2] ![Formula][3] The target quantity of the model is the expected number of HIV transmissions in the study population in transmission direction *g* ⟶ *h* (male-to-female or female-to-male), survey round *r* (R10 to R18) in which infection occurred, and 1-year age bands *i, j* of the source and recipient populations respectively, which we denote by ![Graphic][4]. We considered that the expected number of HIV transmissions in the study population is characterized by transmission risk and modulated by the number of infectious and susceptible individuals, which prompted us to express ![Graphic][5] in the form of a standard discrete-time susceptible-infected (SI) model, ![Formula][6] where ![Graphic][7] is the transmission rate exerted by one infected, virally un-suppressed individual of gender *g* and age *i* on one person in the uninfected (“susceptible”) population of the opposite gender *h* and age *j* in a standardized unit of time in round *r*. With model (2), we express expected transmission flows with a population-level mechanism of how transmission rates from individuals with unsuppressed HIV act on the susceptible population, and we preferred model (2) over a purely phenomenological model of the ![Graphic][8] for the generalizing insights it provides. The main simplifying approximations in (2) are that all quantities on the right-hand side of (2) are in discrete time and constant in each round, meaning we approximate over changes in population size, HIV prevalence, and viral suppression at a temporally finer scale, and assume further that one generation of transmissions occurs from individuals with unsuppressed HIV in each round. Importantly, in this framework, we can then relate the expected transmission flows to the HIV incidence dynamics and the data from the longitudinal incidence cohort by summing in (2) over the sources of infections, ![Formula][9] ![Formula][10] where ![Graphic][11] is the incidence rate per census-eligible, susceptible person of gender *h* and age *j* in round ![Graphic][12] and per unit time ![Graphic][13]. Estimates of ![Graphic][14] were calculated in units of 100 person-years as described above and shown in Fig. 1c, and we will constrain the semi-parametric Poisson flow model using these estimates. From the model output, we are primarily interested in the transmission flows and transmission sources during each round as quantities out of 100%, defined respectively by ![Formula][15] ![Formula][16] ![Formula][17] In words, (4b) quantifies the sources of infection in individuals of gender *h* and age *j* in round *r* such that the sum of ![Graphic][18] over *i* equals one, and (4c) quantifies the sources of infection in the entire population in round *r* that originate from the group of individuals of gender *g* and age *i* such that the sum of ![Graphic][19] over *g* and *i* equals one. The width of the boxplots in Fig. 2b shows (4b) and Fig. 2a, c show (4c). #### Specification of susceptible and infected individuals The number ![Graphic][20] of the susceptible population of gender *h* and age *j* was calculated by multiplying the smoothed estimate ![Graphic][21] of the census-eligible population of gender *h* and age *j* (shown in Extended Data Fig. 1a-b) with 1 minus the posterior median estimate of HIV prevalence ![Graphic][22] in census-eligible individuals of gender *h* and age *j* of round *r* (calculated as described further above). To specify the number ![Graphic][23] of individuals with unsuppressed HIV of gender *g* and age *i*, we multiplied the smoothed estimate ![Graphic][24] of the census-eligible population of gender *g* and age *i* of round *r* (shown in Extended Data Fig. 1a-b) with the posterior median estimate of HIV prevalence in the census-eligible population of gender *g* and age ![Graphic][25] with 1 minus the posterior median estimate ![Graphic][26] of the proportion of census-eligible individuals of gender *g* and age *i* in round *r* that have suppressed HIV (calculated as described further above and shown in Extended Data Fig. 8d). The start and end times of each survey round, ![Graphic][27] and ![Graphic][28] were set as shown in Fig. 1b and specified in units of years, so that the transmission intensity is also expressed in units of years. #### Bayesian model We first present the likelihood of the observed counts of transmission flows ![Graphic][29] under the semi-parametric Poisson flow model that is parameterized in terms of (2). The phylogenetically reconstructed source-recipient pairs capture only a subset of incidence events, and so it is important to characterize the sampling frame. As in Xi, X. *et al*. 29, we consider the unknown transmission events ![Graphic][30] in round *r* and assume these are sampled at random within each strata with probabilities that factorise into sampling probabilities of sources of age *i* and gender *g* and sampling probabilities of recipients of age *j* and gender ![Graphic][31]. Using (4a), we also let ![Graphic][32], where *Z**r* is the total number of infection events in round *r*. Because we have data from both the transmission and incidence cohorts, we are able to constrain the sampling problem the detection probabilities of incidence events. Specifically, setting ![Graphic][33] we let ![Graphic][34] and ![Graphic][35] set the detection probability to the proportion of the expected number of incident cases of gender *h* and age *j* that could be phylogenetically reconstructed in time period *p*, ![Formula][36] for all rounds *r* in the two time periods R10-R15 and R16-R18. We focused in (5) on time periods due to the limited phylogenetic count data. The advantage in constraining the transmission model with the detection probabilities (5) is that the estimates of the transmission model will be consistent with the incidence dynamics that we already estimated with data from the incidence cohort. Re-arranging terms between Binomial and Multinomial models, we obtain ![Formula][37] ![Formula][38] which shows that the sampling probabilities of recipients ![Graphic][39] terms of the detection probability of infection events, weighted by the relative contribution and sampling of source-specific transmission events to the same incidence group. We still need to specify ![Graphic][40] to complete the sampling model. Here, we approximated the sampling probability of sources with the proportion of individuals of age *i* and gender *h* with unsuppressed virus in round *r* that were ever deep-sequenced. Note that the sampling model (6) will alter the posterior mean transmission flows ![Graphic][41] only when the sampling probabilities ![Graphic][42] and ![Graphic][43] differ between age and gender strata in the same round. Extended Data Fig. 5 visualizes our specifications of ![Graphic][44] and ![Graphic][45] shows that the sampling differences between age and gender groups are relatively modest in any given round, which suggests that the adjustments on the inferred transmission flows based on our modelled sampling probabilities will be modest. In the semi-parametric Poisson flow model of Xi, X. *et al*. 29, the sampling model (6) can be analytically integrated out based on standard thinning properties, which in turn allows us to express the likelihood of observing the phylogenetic data with ![Formula][46] ![Formula][47] ![Formula][48] ![Formula][49] where ![Graphic][50] is the posterior median estimate of the log rate of sexual contacts within communities in one year between one person of age *i* and gender *g* and one person of age *j* and gender *h* that we estimated from the sexual behavior data, and the remaining terms quantify the transmission probability per sexual contact on the log scale. The model is designed in such a way that the log sexual contact rates describe a fixed age-specific non-zero mean surface, and the remaining parameters describe age-specific random deviations around the mean surface. With this approach, any inferred deviations in transmission rates relative to sexual contact rates are informed by the phylogenetic data and robust to prior specifications on the random deviations. Specifically, *γ* is the baseline parameter characterizing overall transmission risk per sexual contact, *γ**g* is a gender-specific offset which is set to zero in the female-to-male direction and a real value in male-to-female direction, *γ**r* a round-specific offset which is set to zero for the first survey round 10, and *γ**p* is a time period specific off-set which is set to zero for the first time period. We assume the age-specific structure of transmission rates in terms of the transmitting partners (denoted by *i*) and recipients (denoted by *j*) are similar across similar ages, and so we can exploit regularising prior densities 29 to learn smooth, latent transmission rate surfaces from the sparse data shown in Extended Data Fig. 3. In detail, we modeled the age-specific structure of transmission rates non-parametrically with 2 time-invariant random functions ![Graphic][51] with two-dimensional inputs on the domain [15, 50] × [15, 50] that characterize age-age interactions in transmission risk for each gender, 2 × 8 random functions ![Graphic][52] with one-dimensional inputs that characterize time trends in the age of recipients for each gender for survey rounds after round 10, and 2 random functions ![Graphic][53] with one-dimensional inputs that characterize time trends in the age of transmitting partners for each gender for the second time period. We attach to each of these random functions computationally efficient B-splines projected Gaussian process (GP) priors 89, which we constructed by describing the random functions with cubic B-splines over equidistant knots and modeling the prior relationship of the B-splines parameters with GPs with squared exponential kernels with variance and length-scale hyper-parameters, denoted respectively by *s*2 and *𝓁*. The prior densities of our Bayesian model are ![Formula][54] ![Formula][55] ![Formula][56] ![Formula][57] ![Formula][58] ![Formula][59] ![Formula][60] ![Formula][61] ![Formula][62] where the 2 × 8 recipient-specific time-varying 1D B-splines GPs each have squared exponential kernels with hyper-parameters ![Graphic][63], the 2 source-specific time-varying 1D B-splines GPs each have squared exponential kernels with hyper-parameters ![Graphic][64], and the 2 time-invariant 2D B-splines GPs each have squared exponential kernels with hyper-parameters ![Graphic][65] and ![Graphic][66] decomposed as follows, ![Formula][67] We constrain the model further with a pseudo-likelihood term so that the model’s implied incidence rate ![Graphic][68] in (3b) is around the MLE incidence rate estimate obtained from the incidence cohort. We took this approach in lieu of fitting the model to both the source-recipient and individual-level incidence exposure data to bypass extreme computational runtimes 12, and in the context that the source-recipient data are not informative of incidence dynamics 90. Specifically, we fitted log-normal distributions to the 1, 000 × 50 Monte Carlo replicate rate estimates for individuals of gender *h* and age *j* in round *r* (see above) using the lognorm R package version 0.1.691, and then set ![Formula][69] where ![Graphic][70] and ![Graphic][71] denote respectively the parameters of the fitted log-normal distributions, and the left-hand side is calculated from (7b) and matches the model’s incidence rate ![Graphic][72] in (3b). #### Computational inference Model (7-10) was fitted with Rstan version 2.21.0, using Stan’s adaptive HMC sampler 63 with 4 chains for 3,500 iterations including 500 warm-up iterations. Convergence and mixing were good, with highest Rhat value of 1.0027 and lowest effective sample sizes of 1444. The model presented the data well, with 99.63% data point inside 95% posterior predictive intervals and the fitted model was consistent with all the available data (Extended Data Fig. 6), indicating that the data met the assumptions of the statistical model. There were no divergent transitions, suggesting non-pathological posterior topologies. ### Counterfactual interventions We investigated —given the inferred transmission flows— the hypothetical impact of targeted counterfactual intervention scenarios *c* on predicted incidence reductions in women in the most recent survey round 18. In the model, counterfactual interventions were implemented by calculating the expected number of transmission flows (2) into women under counterfactual *c* that fewer men of age *i* had remained with un-suppressed HIV in survey round 18, which we denote by ![Graphic][73]. We obtained the expected number of incident cases in women of age *j* in round 18 in counterfactual *c* via ![Formula][74] where uncertainty in the posterior age-specific transmission rates after fitting model (7-10) is integrated out. The predicted incidence rate reductions were based on comparing the counterfactuals 1) to the red cases in women in the corresponding age group (3b),![Graphic][75]. #### Closing half the gap in viral suppression rates in men relative to women In this scenario, we considered the impact of reducing by half the gap in the proportion of men with unsuppressed HIV compared to the same proportion in women. To this end, we first calculated for each 1-year age band the average of the estimated proportion of census-eligible infected men in round 18 with suppressed virus and the same proportion in women, ![Graphic][76]. Next, we set ![Graphic][77] to the smoothed estimate of census-eligible men of age *i* in round R18 multiplied with the posterior median estimate of HIV prevalence in census-eligible men of age *i*, and with ![Graphic][78]. #### Closing the gap in viral suppression rates in men relative to women In this scenario, we considered the impact of achieving the same proportions of men with unsuppressed HIV as in women. To this end, we set ![Graphic][79] to the smoothed estimate of census-eligible men of age *i* in round R18 multiplied with the posterior median estimate of HIV prevalence in census-eligible men of age *i*, and with ![Graphic][80]. #### 95-95-95 in men In this scenario, we considered the impact of achieving viral suppression in 85.7% (0.95 × 0.95 × 0.95) in each 1-year age group of men with HIV. The number of remaining men with unsuppressed HIV in round 18, ![Graphic][81], was calculated by multiplying the smoothed estimate of the census-eligible men of age *i* in round R18 with the posterior median estimate of HIV prevalence in the census-eligible men of age *i*, and with 1 − 0.857. ### Sensitivity analyses #### Sensitivity in incidence rate estimates to the GAM incidence model specification The longitudinal age-specific HIV incidence rates of the central analysis were estimated with a log-link generalized additive effects Poisson regression model with a linear predictor comprising relatively simple main and interaction effects by age and survey round, fitted to individual-level 0/1 incidence outcomes and exposure times specified as offset on the log scale. To assess sensitivity against the relatively simple linear predictor, we considered a more complex mean specification comprising in-dependent LOESS smoothers to capture age-specific incidence trends in each survey round, and fitted this mean model for computational reasons to crude HIV incidence rates. Specifically, we fitted LOESS regressions as implemented in the R package stats version 3.6.2 with span argument set to 0.7 to the crude age-, gender- and round-specific HIV incidence rates in all 50 imputation data sets, and weighted by the corresponding, group-level aggregated exposure times. The HIV incidence rate estimates under the LOESS model had as expected a smaller mean absolute error against the crude estimates as compared against the GAM model (0.0048 [0.0046-0.0051] versus 0.0053 [0.0051-0.0056]) (Supplementary Fig. S3). Overall, the contribution of men to incidence was more variable across rounds while the shifts in the median age at infection were similar in the central and this sensitivity analysis (Supplementary Table S10). #### Sensitivity in incidence rate and transmission flow estimates to limited communities Over time some communities were added and others left the RCCS (see Supplementary Table S2). We repeated our analysis on the subset of 28 consecutively surveyed communities. We found similar incidence rates with slightly faster declines in male new infections and larger gender disparities (Supplementary Fig. S4). All other primary findings remained insensitive (Supplementary Table S10). #### Sensitivity in estimating transmission flows to uncertainty in infection time estimates In the central analysis, phyloTSI infection time estimates associated to source-recipient pairs were refined using the inferred transmission direction, age, and serohistory data. To assess sensitivity to the infection time estimates used, we inferred transmission flows on the basis of the raw phyloTSI infection time estimates as long as they were compatible with the inferred transmission direction, and otherwise on the basis of the refined estimates. Overall, we found source-recipient pairs were potentially allocated to earlier or later time periods reflecting the wide uncertainty in infection time estimates, though across the sample the age distribution of sources and recipients was remarkably stable (Extended Data Fig. 4). All primary findings were insensitive to using the raw infection time estimates (Supplementary Table S10). #### Sensitivity in time since infection estimates to higher transmissibility during acute infection In the central analysis, transmission flows were estimated using the centre of gravity of the uncertainty region associated with the refined infection time estimates. To account for higher transmission rates during acute infection of the transmitting partner 88, we assumed that the transmission hazard was 5 times higher in the first 2 months after infection of the transmitting partner as compared to the following period, and obtained the resulting mean infection time estimate under this assumption by generalizing our Monte Carlo approach used in the central analysis to an importance sampling approach under piecewise linear transmission hazards. The primary results were insensitive to these changes as less than 5% of source-recipient pairs were attributed to different survey rounds (Supplementary Table S10). #### Sensitivity in estimating transmission flows to right censoring of likely transmission pairs The RCCS transmission cohort was defined retrospectively and so it is possible that some transmission events, especially in later rounds, remain as of yet unseen because the corresponding individuals are not yet in the survey or do not yet have virus deep-sequenced. To assess sensitivity to right censoring, we excluded from analysis those source-recipient pairs for which virus of the source or the recipient was deep-sequenced only after rounds 17, 16 and 15. The primary findings were insensitive to these analyses because the probabilities of detecting infection events in the phylogenetic data changed accordingly (Supplementary Table S10 and Supplementary Fig. S5). #### Sensitivity in estimated transmission flows to limited sample size of likely transmission pairs The number of observed infection events in the incidence cohort was ≈4 times larger than the number of reconstructed transmission events, prompting us to explore the effect of sampling uncertainty on the transmission flow estimates. We bootstrap sampled source-recipient pairs at random with replacement three times, and repeated inferences on these bootstrap samples. Our primary findings remained insensitive (Supplementary Table S10). #### Sensitivity in estimated transmission flows to modelled sampling estimates The sampling adjustments in (6) require assumptions including that sampling is independent of infection and transmission, independent between source and recipient, at random within strata, and well approximated by approximating sources with individuals with unsuppressed virus. We repeated flow inferences without any adjustments and without adjustments for potentially unequal sampling of sources. Our primary findings were insensitive across these analyses (Supplementary Table S10). #### Sensitivity in transmission flow estimates to the phylo-SI model specification In the central analysis, the log transmission rates that underpin the estimated transmission flows were estimated using the linear predictor in (7c), and this model specification was associated with overall smallest mean absolute error and posterior predictive coverage as shown in Supplementary Table S6 against the following alternative models, ![Formula][82] ![Formula][83] ![Formula][84] ![Formula][85] ![Formula][86] ![Formula][87] Models specifying transmission rates without a round-specific random function on the age of infected individuals, (12a)-(12c), did not fit the data well (Supplementary Table S6). The remaining models, (12d)-(12f) performed as well as the model used in the central analysis (Supplementary Table S6) and our primary findings remained insensitive (Supplementary Table S10). #### Sensitivity in counterfactual intervention impacts to assumptions on viral suppression levels in non-participants Infection and viremia in the non-participant census-eligible population remained unknown and in the central analysis, we considered as proxy of virus suppression levels among non-participants data from first-time participants. We performed two sensitivity analyses, assuming first that all non-participants with HIV were also viremic across all rounds, and assuming second that virus suppression was identical among non-participants and participants of the same age, gender and survey round. Together, the two scenarios likely encompass the true, unknown viral suppression levels in non-participants. These scenarios were implemented by updating the number of individuals with viremia in (2), and refitting the model. The sensitivity analysis assuming all non-participants with HIV were viremic resulted in larger predicted incidence reductions in women around 75%, while the sensitivity analysis assuming virus suppression was the same among non-participants as among participants of the same age, gender and survey round resulted in similar predicted incidence reductions in women than in the central analysis (Supplementary Table S10). #### Sensitivity in counterfactual intervention impacts to potentially higher HIV prevalence in non-participants In the central analysis, we assumed that HIV prevalence was the same in participants and non-participants of the same age, gender and survey round. We considered three sensitivity analyses, assuming first that prevalence was 25% higher in male non-participants compared to male participants of the same age, gender and survey round, assuming second that prevalence was 25% higher in female non-participants compared to female participants of the same age, gender and survey round, and assuming third that prevalence was 25% higher in female and male non-participants compared to female and male participants of the same age, gender and survey round respectively. These scenarios were implemented by updating the number of virally unsuppressed individuals in (2), and refitting the model. Our primary findings remained insensitive (Supplementary Table S10). #### Sensitivity in counterfactual intervention impacts to lower viral suppression thresholds Different definitions of HIV suppression are currently operational, and we considered the effect of lower thresholds to define viral suppression (<200 copies/mL) than in the central analysis (<1,000 copies/mL). This scenario was implemented by re-estimating the age- and gender-specific proportions of individuals with HIV in the study population who had suppressed virus at the lower threshold, re-calculating gaps in viral suppression levels in men relative to women, and re-calculating the additional number of men needed to reach and maintain viral suppression in the counterfactual intervention scenarios. We found slightly smaller gender gaps in viral suppression at the lower threshold and the predicted incidence reduction in women in the counterfactual that assessed closing the suppression gap in men was around 45%, and all other findings remained insensitive (Supplementary Table S10). ## Data Availability Pseudo-anonymised data from the RCCS incidence and transmission cohort as well as pseudo-anonymised deep-sequence phylogenies to reproduce all analyses are available from Zenodo ([https://zenodo.org/record/8412741](https://zenodo.org/record/8412741)) as open-access data set under the CC-BY-4.0 license 92. HIV consensus sequences are available from Zenodo ([https://zenodo.org/records/10075815](https://zenodo.org/records/10075815)) and the PANGEA-HIV sequence repository ([https://github.com/PANGEA-HIV/PANGEA-Sequences](https://github.com/PANGEA-HIV/PANGEA-Sequences)) as open-access data set under the CC-BY-4.0 license, with identifiers changed to ensure participants cannot be identified from this data set. Additional deep-sequence HIV-1 reads can be requested from PANGEA-HIV under a managed access policy due to privacy and ethical reasons, which aligns with UNAIDS ethical guidelines. The process for accessing data, the PANGEA-HIV Data Sharing Policy and a detailed description of what data are available is laid out in full at ([https://www.pangea-hiv.org/join-us](https://www.pangea-hiv.org/join-us)). Briefly, applicants can apply to receive additional data by submitting a concept sheet proposal in which they explain the research question and how they will mitigate potential risks to participant privacy. In line with requirements for PANGEA members, applicants will be asked to present proof of human subject research training and comply with PANGEA-internal publication agreements. PANGEA encourages external applicants to collaborate with the researchers who generated the data. For more information contact PANGEA project manager Lucie Abeler-Dörner (lucie.abeler-dorner@bdi.ox.ac.uk). The time frame for a response to requests is 2-4 weeks. Additional cohort data can be requested from RHSP. Because HIV transmission is criminalized in Uganda and due to further privacy considerations, RHSP maintains a controlled access data policy for corresponding epidemiological metadata and corresponding data collection tools. In brief, RHSP policy requires individuals to submit an RSHSP data request form (available upon request from info{at}rhsp.org or gkigozi{at}rhsp.org) and a brief concept note (1-2 pages) detailing their research questions and methods. In addition, researchers are asked to provide a curriculum vitae/resume along with proof of human subjects research training. Concept sheets can be submitted to Dr. Godfrey Kigozi (gkigozi{at}rhsp.org), executive director of the RHSP. Only individuals named on the original data request and who provide the request, CV/resume and HSR training, are permitted access to the data. Released data are not to be reused for other purposes outside of approved concepts. The time frame for a response to requests is 2-4 weeks. ## Code availability Code to reproduce all analyses is freely available on GitHub version 1.1.2 under the GNU General Public License version 3.0 at the repository ([https://github.com/MLGlobalHealth/phyloSI-RakaiAgeGender](https://github.com/MLGlobalHealth/phyloSI-RakaiAgeGender)). ## Author contributions OR and MKG designed the study. OR, MKG, JK, PGF, DS oversaw the study. RG, RS, EK, VS, LAD, DB, LWC, CF, TG, RHG, JJ, GK, LM, OL, TQ, SJR, JSa, NKS, JSs, LT, MJW, DS, JK, MKG oversaw and performed data collection. MM, ABr, XX, EK, VS, AA, ABl, YC, SD, TG, MH, SS, LT, MKG, OR contributed to the analysis. MM, ABr, XX, ABl, YC, MKG, OR wrote the first draft. None supplied materials. All authors discussed the results and contributed to the revision of the final manuscript. ## Competing interests OR, MKG, CF, DB report grants from the Bill & Melinda Gates Foundation during the conduct of this study. MKG, MJW, RHG, LCW report grants from the National Institutes of Health during the conduct of this study. ABr, YC, XX report an EP-SRC PhD studentship during the conduct of this study. Dr. Wawer and Dr. Gray are paid consultants to the Rakai Health Sciences Program and serves on its Board of Directors. This arrangement has been reviewed and approved by the Johns Hopkins University in accordance with its conflict of interest policies. ## Rakai Health Sciences Program consortium authors Larry W Chang3,2,11, Ronald M Galiwango2, M Kate Grabowski2,13,11, Ronald H Gray11, Jade C Jackson13, Joseph Kagaayi2, Edward Nelson Kankaka3,4, Godfrey Kigozi2, Oliver Laeyendecker14,15, Thomas C Quinn13,14,15, Steven J. Reynolds2,14,15, John Santelli17, David Serwadda2,18, Nelson K. Sewankambo18, Joseph Ssekasanvu11, Robert Ssekubugu2, Victor Ssempijja5,6, Maria J Wawer2,11 ## PANGEA-HIV consortium authors Lucie Abeler Dörner7, David Bonsall9,10, Christophe Fraser10,7, Tanya Golubchik12,7, M Kate Grabowski2,13,11, Joseph Kagaayi2, Thomas C Quinn13,14,15, Oliver Ratmann1, Maria J Wawer2,11 ## S1 Supplementary Figures ![](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F5/graphic-39.medium.gif) [](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F5/graphic-39) ![](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F5/graphic-40.medium.gif) [](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F5/graphic-40) ![](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F5/graphic-41.medium.gif) [](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F5/graphic-41) ![](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F5/graphic-42.medium.gif) [](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F5/graphic-42) ![](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F5/graphic-43.medium.gif) [](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F5/graphic-43) Supplementary Fig. S1: Flowchart of census eligible individuals through to individuals in the incidence and transmission cohorts. ![Supplementary Fig. S2:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F6.medium.gif) [Supplementary Fig. S2:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F6) Supplementary Fig. S2: Schema illustrating the refinement of phylogenetic time since infection estimates. ![Supplementary Fig. S3:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F7.medium.gif) [Supplementary Fig. S3:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F7) Supplementary Fig. S3: Comparison of incidence rate estimates under an individual-level additive effects Poisson regression model and a population-level LOESS model with independent age effects in each survey round. (**a**) Mean and 95% uncertainty ranges of longitudinal age-specific incidence rates obtained with the individual-level additive effects Poisson regression model used in the central analysis (**b**) Same using a population-level LOESS model with independent age effects in each survey round. ![Supplementary Fig. S4:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F8.medium.gif) [Supplementary Fig. S4:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F8) Supplementary Fig. S4: Comparison of incidence rate estimated on data containing all communities and data subset to 28 continuously surveyed communities (**a**) Mean and 95% uncertainty ranges of longitudinal age-specific incidence rates estimated on data from all communities surveyed (**b**) Same using data subset to 28 continuously surveyed communities. ![Supplementary Fig. S5:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2023/11/13/2023.03.16.23287351/F9.medium.gif) [Supplementary Fig. S5:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/F9) Supplementary Fig. S5: Sensitivity in estimating the age of transmitting partners to right censoring of likely transmission pairs. Posterior median (line) and 95% credible interval (ribbon) of the age of male transmitting partners by the age of the infected female (x-axis) by survey round (row facet) for the central and sensitivity analyses (column facet). Median and 95% credible interval of the age of male transmitting partners across the age of the infected female is indicated with a triangle and an error bar. ## S2 Supplementary Tables View this table: [Supplementary Table S1:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T2) Supplementary Table S1: Characteristics of the RCCS study population. View this table: [Supplementary Table S2:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T3) Supplementary Table S2: Communities surveyed by RCCS in rounds 10-18. View this table: [Supplementary Table S3:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T4) Supplementary Table S3: Characteristics of the longitudinal HIV incidence cohort. View this table: [Supplementary Table S4:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T5) Supplementary Table S4: Model comparison for estimating longitudinal, age-specific incidence rates. View this table: [Supplementary Table S5:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T6) Supplementary Table S5: Longitudinal HIV deep-sequencing. View this table: [Supplementary Table S6:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T7) Supplementary Table S6: Model comparison for estimating longitudinal, age-specific transmission flows. View this table: [Supplementary Table S7:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T8) Supplementary Table S7: Longitudinal HIV transmission flows by age and gender. View this table: [Supplementary Table S8:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T9) Supplementary Table S8: Sexual behaviour characteristics in RCCS participants, round 15, October 08 2011 - July 05 2013. View this table: [Supplementary Table S9:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T10) Supplementary Table S9: Self-reported ART use and viral suppression in RCCS participants with HIV. View this table: [Supplementary Table S10:](http://medrxiv.org/content/early/2023/11/13/2023.03.16.23287351/T11) Supplementary Table S10: Sensitivity analyses. ## Acknowledgements We thank all contributors, program staff and participants to the Rakai Community Cohort Study; all members of the PANGEA-HIV consortium, the Rakai Health Sciences Program, and CDC Uganda for comments on an earlier version of the manuscript; the Imperial College Research Computing Service ([https://doi.org/10.14469/hpc/2232](https://doi.org/10.14469/hpc/2232)) and the Biomedical Research Computing Cluster at the University of Oxford for providing the computational resources to perform this study; the Office of Cyberinfrastructure and Computational Biology at the National Institute for Allergy and Infectious Diseases for data management support; and Zulip for sponsoring team communications through the Zulip Cloud Standard chat app. This study was supported by the Bill & Melinda Gates Foundation (OPP1175094 to CF, OPP1084362 to Prof Deenan Pillay); the National Institute of Allergy and Infectious Diseases (U01AI051171 to RHG, U01AI075115 to RHG, UM1AI069530-16 to Dr Mary Glenn Fowler, Dr Philippa Musoke, and Dr Aaron Tobian, R01AI087409 to RHG, U01AI100031 to RHG, R01AI110324 to RHG, R01AI114438 to MJW, K25AI114461 to Dr Xiangrong Kong, R01AI123002 to Dr Cindy Liu, K01AI125086 to MKG, R01AI128779 to Dr Aaron Tobian, R01AI143333 to LWC, R21AI145682 to Dr Caitlin Kennedy, R01AI155080 to MKG, ZIAAI001040 to TCQ); the National Institute of Mental Health (F31MH095649 to Dr Jennifer Wagman, R01MH099733 to Ned Sacktor and MJW, R01MH107275 to LWC, R01MH115799 to MJW and LWC, U19MH110001 to Dr Mary McKay and Dr Fred Ssewamala); the National Institute of Child Health and Development (R01HD038883 to RHG, R01HD050180 to MJW, R01HD070769 to MJW, R01HD091003 to JSa); the Division of Intramural Research of the National Institute for Allergy and Infectious Diseases (TCQ, OL and SJR); NIAID K01AA024068 to Dr Jennifer Wagman; the National Heart, Lung, and Blood Institute (R01HL152813 to LWC); the Fogarty International Center (D43TW009578 to RHG, D43TW010557 to LWC); the Doris Duke Charitable Foundation to Dr Aaron Tobian; the Johns Hopkins University Center for AIDS Research (P30AI094189 to Dr Richard Chaisson); the U.S. President’s Emergency Plan for AIDS Relief (PEP-FAR) through the Centers for Disease Control and Prevention (NU2GGH000817 to RHSP); the Engineering and Physical Sciences Research Council through the EPSRC Centre for Doctoral Training in Modern Statistics and Statistical Machine Learning at Imperial College London and Oxford University (EP/S023151/1 to Prof Axel Gandy); and the Imperial College London President’s PhD Scholarship fund to YC. The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript. The findings and conclusions in this report are those of the author(s) and do not necessarily represent the official position of the Centers for Disease Control and Prevention. ## Footnotes * Updated data availability statement and acknowledgments * Received March 16, 2023. * Revision received November 13, 2023. * Accepted November 13, 2023. * © 2023, Posted by Cold Spring Harbor Laboratory This pre-print is available under a Creative Commons License (Attribution 4.0 International), CC BY 4.0, as described at [http://creativecommons.org/licenses/by/4.0/](http://creativecommons.org/licenses/by/4.0/) ## References 1. [1].UNAIDS. 2021 UNAIDS global AIDS update — confronting in-equalities — lessons for pandemic responses from 40 years of AIDS (2021). Available at [https://www.unaids.org/sites/default/files/media\_asset/2021-global-aids-update_en.pdf](https://www.unaids.org/sites/default/files/media_asset/2021-global-aids-update_en.pdf). 2. [2].UNAIDS. Women and HIV — a spotlight on adolescent girls and young women (2019). Available at [https://www.unaids.org/sites/default/files/media\_asset/2019\_women-and-hiv\_en.pdf](https://www.unaids.org/sites/default/files/media_asset/2019_women-and-hiv_en.pdf). 3. [3].Goga, A. et al. Centring adolescent girls and young women in the HIV and COVID-19 responses. Lancet 396, 1864–1866 (2020). 4. [4].UNAIDS. Dangerous inequalities: world AIDS day report 2022 (2022). Available at [https://www.unaids.org/sites/default/files/media\_asset/dangerous-inequalities_en.pdf](https://www.unaids.org/sites/default/files/media_asset/dangerous-inequalities_en.pdf). 5. [5].UNAIDS. Full report — in danger: UNAIDS global AIDS update 2022 (2022). Available at [https://www.unaids.org/sites/default/files/media\_asset/2022-global-aids-update_en.pdf](https://www.unaids.org/sites/default/files/media_asset/2022-global-aids-update_en.pdf). 6. [6].Gupta, R. K. et al. HIV-1 drug resistance before initiation or re-initiation of first-line antiretroviral therapy in low-income and middle-income countries: a systematic review and meta-regression analysis. Lancet Infectious Diseases 18, 346–355 (2018). 7. [7].World Health Organization. HIV drug resistance report 2021 (2021). Available at [https://www.who.int/publications-detail-redirect/9789240038608](https://www.who.int/publications-detail-redirect/9789240038608). 8. [8].UNAIDS. Resources and financing for the AIDS response (2022). Available at [https://www.unaids.org/en/topic/resources](https://www.unaids.org/en/topic/resources). 9. [9].UNAIDS. Agenda item 2: Report of the 48th PCB meeting (2021). Available at [https://www.unaids.org/sites/default/files/media\_asset/PCBSS\_Report\_48th\_PCB\_EN\_rev3.pdf](https://www.unaids.org/sites/default/files/media\_asset/PCBSS\_Report\_48th_PCB_EN_rev3.pdf). 10. [10].United States Department of State. Deams partnership (2022). Available at [https://www.state.gov/wp-content/uploads/2022/09/PEPFAR-Strategic-Direction_FINAL.pdf](https://www.state.gov/wp-content/uploads/2022/09/PEPFAR-Strategic-Direction_FINAL.pdf). 11. [11].United States Department of State. Reimagining PEPFAR at 20 to end the HIV/AIDS pandemic by 2030 (2022). Available at [https://www.state.gov/reimagining-pepfar-at-20-to-end-the-hiv-aids-pandemic-by-2030/](https://www.state.gov/reimagining-pepfar-at-20-to-end-the-hiv-aids-pandemic-by-2030/). 12. [12].Risher, K. A. et al. Age patterns of HIV incidence in eastern and southern Africa: a modelling analysis of observational population-based cohort studies. Lancet HIV 8, e429–e439 (2021). 13. [13].Akullian, A. et al. Large age shifts in HIV-1 incidence patterns in KwaZulu-Natal, South Africa. Proceedings of the National Academy of Sciences of the United States of America 118, e2013164118 (2021). [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxODoiMTE4LzI4L2UyMDEzMTY0MTE4IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjMvMTEvMTMvMjAyMy4wMy4xNi4yMzI4NzM1MS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 14. [14].Grabowski, M. K. et al. HIV prevention efforts and incidence of HIV in Uganda. New England Journal of Medicine 377, 2154–2166 (2017). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1056/NEJMoa1702150&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=29171817&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 15. [15].Vandormael, A., Akullian, A., Siedner, M., de Oliveira, T., Bärnighausen, T. & Tanser, F. Declines in HIV incidence among men and women in a South African population-based cohort. Nature communications 10, 1–10 (2019). 16. [16].Grabowski, M. K. et al. The role of viral introductions in sustaining community-based HIV epidemics in rural Uganda: evidence from spatial clus-tering, phylogenetics, and egocentric transmission models. PLoS medicine 11, e1001610 (2014). 17. [17].Joshi, K. et al. Declining HIV incidence in sub-Saharan Africa: a systematic review and meta-analysis of empiric data. Journal of the International AIDS Society 24, e25818 (2021). 18. [18].Chang, L. W. et al. Heterogeneity of the HIV epidemic in agrarian, trading, and fishing communities in Rakai, Uganda: an observational epidemiological study. Lancet HIV 3, e388–e396 (2016). 19. [19].Abeler-Dörner, L., Grabowski, M. K., Rambaut, A., Pillay, D. & Fraser, C. PANGEA-HIV 2: phylogenetics and networks for generalised epidemics in africa. Current Opinion in HIV and AIDS 14, 173 (2019). 20. [20].Hall, M. et al. Demographic characteristics of sources of HIV-1 transmission in the era of test and treat. medRxiv (2022). Preprint at [https://www.medrxiv.org/content/early/2022/10/13/2021.10.04.21263560.full.pdf](https://www.medrxiv.org/content/early/2022/10/13/2021.10.04.21263560.full.pdf). 21. [21].Vandormael, A., Akullian, A., Siedner, M., de Oliveira, T., Bärnighausen, T. & Tanser, F. Declines in HIV incidence among men and women in a South African population-based cohort. Nature Communications 10, 5482 (2019). 22. [22].Ratmann, O. et al. Inferring HIV-1 transmission networks and sources of epidemic spread in Africa with deep-sequence phylogenetic analysis. Nature Communications 10, 1411 (2019). 23. [23].Fisher, M. et al. Determinants of HIV-1 transmission in men who have sex with men: a combined clinical, epidemiological and phylogenetic approach. Aids 24, 1739–1747 (2010). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1097/QAD.0b013e32833ac9e6&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=20588173&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000279396600015&link_type=ISI) 24. [24].Ratmann, O. et al. Sources of HIV infection among men having sex with men and implications for prevention. Science translational medicine 8, 320ra2 (2016). [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MTE6InNjaXRyYW5zbWVkIjtzOjU6InJlc2lkIjtzOjEyOiI4LzMyMC8zMjByYTIiO3M6NDoiYXRvbSI7czo1MDoiL21lZHJ4aXYvZWFybHkvMjAyMy8xMS8xMy8yMDIzLjAzLjE2LjIzMjg3MzUxLmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 25. [25].Poon, A. F. et al. Near real-time monitoring of HIV transmission hotspots from routine HIV genotyping: an implementation case study. Lancet HIV 3, e231–e238 (2016). 26. [26].Wymant, C. et al. PHYLOSCANNER: inferring transmission from within-and between-host pathogen genetic diversity. Molecular Biology and Evolution 35, 719–733 (2018). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/molbev/msx304&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=29186559&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 27. [27].Golubchik, T. et al. HIV-phyloTSI: Subtype-independent estimation of time since HIV-1 infection for cross-sectional measures of population incidence using deep sequence data. medRxiv (2022). Preprint at [https://www.medrxiv.org/content/early/2022/05/16/2022.05.15.22275117.full.pdf](https://www.medrxiv.org/content/early/2022/05/16/2022.05.15.22275117.full.pdf). 28. [28].Ratmann, O. et al. Quantifying HIV transmission flow between high-prevalence hotspots and surrounding communities: a population-based study in Rakai, Uganda. Lancet HIV 7, e173–e183 (2020). 29. [29].Xi, X. et al. Inferring the sources of HIV infection in Africa from deep-sequence data with semi-parametric Bayesian Poisson flow models. Journal of the Royal Statistical Society: Series C (Applied Statistics) 71, 517–540 (2022). 30. [30].Wilson, D. & Halperin, D. T. “Know your epidemic, know your response”: a useful approach, if we get it right. Lancet 372, 423–426 (2008). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/S0140-6736(08)60883-1&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=18687462&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000258299000005&link_type=ISI) 31. [31].Oliveira, T. d. et al. Transmission networks and risk of HIV infection in KwaZulu-Natal, South Africa: a community-wide phylogenetic study. Lancet HIV 4, e41–e50 (2017). 32. [32].Evans, M. et al. Age-disparate sex and HIV risk for young women from 2002 to 2012 in South Africa. Journal of the International AIDS Society 19, 21310 (2016). 33. [33].Akullian, A., Bershteyn, A., Klein, D., Vandormael, A., Bärnighausen, T. & Tanser, F. Sexual partnership age pairings and risk of HIV acquisition in rural South Africa. AIDS (London, England) 31, 1755–1764 (2017). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1097/QAD.0000000000001553&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 34. [34].Kyle, I. Population level HIV viral load varies by gender, age, and location in Rakai, Uganda (2020). Available at [https://www.croiconference.org/wp-content/uploads/sites/2/posters/2020/1430\_3\_Quinn\_00865.pdf](https://www.croiconference.org/wp-content/uploads/sites/2/posters/2020/1430_3_Quinn_00865.pdf). 35. [35].World Health Organization. Consolidated guidelines on the use of antiretroviral drugs for treating and preventing HIV infection: recommendations for a public health approach (World Health Organization, 2016). 36. [36].Tanser, F. et al. Effect of population viral load on prospective HIV incidence in a hyperendemic rural African community. Science Translational Medicine 9, eaam8012 (2017). [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MTE6InNjaXRyYW5zbWVkIjtzOjU6InJlc2lkIjtzOjE0OiI5LzQyMC9lYWFtODAxMiI7czo0OiJhdG9tIjtzOjUwOiIvbWVkcnhpdi9lYXJseS8yMDIzLzExLzEzLzIwMjMuMDMuMTYuMjMyODczNTEuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 37. [37].Donnelly, C. et al. Gender difference in HIV-1 RNA viral loads. HIV medicine 6, 170–178 (2005). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1111/j.1468-1293.2005.00285.x&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=15876283&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000229374200005&link_type=ISI) 38. [38].Reniers, G., Armbruster, B. & Lucas, A. Sexual networks, partnership mixing, and the female-to-male ratio of HIV infections in generalized epidemics: An agent-based simulation study. Demographic research 33, 425–450 (2015). 39. [39].Quinn, T. C. & Overbaugh, J. HIV/AIDS in women: an expanding epidemic. Science 308, 1582–1583 (2005). [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6Mzoic2NpIjtzOjU6InJlc2lkIjtzOjEzOiIzMDgvNTcyOC8xNTgyIjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjMvMTEvMTMvMjAyMy4wMy4xNi4yMzI4NzM1MS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 40. [40].Glynn, J. R. et al. Why do young women have a much higher prevalence of HIV than young men? A study in Kisumu, Kenya and Ndola, Zambia. Aids 15, S51–S60 (2001). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1097/00002030-200108004-00006&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=11686466&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000171519300006&link_type=ISI) 41. [41].Loevinsohn, G. et al. Effectiveness of voluntary medical male circumcision for human immunodeficiency virus prevention in Rakai, Uganda. Clinical Infectious Diseases 73, e1946–e1953 (2021). 42. [42].Rodger, A. J. et al. Risk of HIV transmission through condomless sex in serodifferent gay couples with the HIV-positive partner taking suppressive an-tiretroviral therapy (PARTNER): final results of a multicentre, prospective, observational study. Lancet 393, 2428–2438 (2019). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/S0140-6736(19)30418-0&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=31056293&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 43. [43].Cohen, M. S. et al. Antiretroviral therapy for the prevention of HIV-1 transmission. New England Journal of Medicine 375, 830–839 (2016). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1056/NEJMoa1600693&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=27424812&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 44. [44].UNAIDS. Global AIDS strategy 2021-2026 - end inequalities end AIDS. (2021). Available at [https://www.unaids.org/sites/default/files/media\_asset/global-AIDS-strategy-2021-2026_en.pdf](https://www.unaids.org/sites/default/files/media_asset/global-AIDS-strategy-2021-2026_en.pdf). 45. [45].Stover, J. et al. Modeling the epidemiological impact of the UNAIDS 2025 targets to end AIDS as a public health threat by 2030. PLOS Medicine 18, e1003831 (2021). 46. [46].Ssempijja, V. et al. High rates of pre-exposure prophylaxis eligibility and associated HIV incidence in a population with a generalized HIV epidemic in Rakai, Uganda. Journal of acquired immune deficiency syndromes (1999) 90, 291–299 (2022). 47. [47].Godfrey-Faussett, P., Frescura, L., Karim, Q. A., Clayton, M., Ghys, P. D. & On Behalf of the 2025 Prevention Targets Working. HIV prevention for the next decade: Appropriate, person-centred, prioritised, effective, combination prevention. PLOS Medicine 19, e1004102 (2022). 48. [48].The Lancet HIV. Addressing inequalities still key to ending HIV/AIDS. Lancet HIV 10, e1 (2023). 49. [49].Havlir, D. et al. What do the universal test and treat trials tell us about the path to HIV epidemic control? Journal of the International AIDS Society 23, e25455 (2020). 50. [50].The PHIA Project. PHIA data manager (2022). Available at [https://phia-data.icap.columbia.edu/](https://phia-data.icap.columbia.edu/). 51. [51].Magosi, L. E. et al. Deep-sequence phylogenetics to quantify patterns of HIV transmission in the context of a universal testing and treatment trial - BCPP/Ya Tsie trial. eLife 11, e72657 (2022). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.7554/eLife.72657&link_type=DOI) 52. [52].The RISE consortium. Engaging men in HIV testing, linkage, and retention in care (2020). Available at [https://www.jhpiego.org/wp-content/uploads/2021/01/8-RISE-Engaging-Men-Brief-1.pdf](https://www.jhpiego.org/wp-content/uploads/2021/01/8-RISE-Engaging-Men-Brief-1.pdf). 53. [53].Colvin, C. J. Strategies for engaging men in HIV services. Lancet HIV 6, e191–e200 (2019). 54. [54].Sithole, N. et al. Implementation of HIV self-testing to reach men in rural uMkhanyakude, KwaZulu-Natal, South Africa. A DO-ART trial sub study. Frontiers in Public Health 9, 652887 (2021). 55. [55].United States Department of State. The United States president’s emergency plan for AIDS relief (2022). Available at [https://www.state.gov/pepfar/](https://www.state.gov/pepfar/). 56. [56].Farquhar, C., Masyuko, S. & Mugo, P. Social network–based strategies to improve uptake of HIV testing and linkage to care among men who have sex with men in Sub-Saharan Africa. JAMA Network Open 5, e220155 (2022). 57. [57].De Cock, K. M., Barker, J. L., Baggaley, R. & El Sadr, W. M. Where are the positives? HIV testing in sub-Saharan Africa in the era of test and treat. AIDS 33, 349 (2019). 58. [58].Mugavero, M. J. et al. Beyond core indicators of retention in HIV care: missed clinic visits are independently associated with all-cause mortality. Clinical Infectious Diseases 59, 1471–1479 (2014). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/cid/ciu603&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=25091306&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 59. [59].Mukumbang, F. C. Leaving no man behind: how differentiated service delivery models increase men’s engagement in HIV care. International Journal of Health Policy and Management 10, 129–140 (2021). 60. [60].Kripke, K., Eakle, R., Cheng, A., Rana, S., Torjesen, K. & Stover, J. The case for prevention – Primary HIV prevention in the era of universal test and treat: A mathematical modeling study. eClinicalMedicine 46, 101347 (2022). 61. [61].Rosenberg, N. E. et al. Adult HIV-1 incidence across 15 high-burden countries in sub-saharan africa from 2015 to 2019: a pooled analysis of nationally Changing drivers of HIV infection in Africa 45 representative data. The Lancet HIV 10, e175–e185 (2023). 62. [62].Galiwango, R. M. et al. Evaluation of current rapid HIV test algorithms in Rakai, Uganda. Journal of Virological Methods 192, 25–27 (2013). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1016/j.jviromet.2013.04.003&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=23583487&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 63. [63].Carpenter, B. et al. Stan: A probabilistic programming language. Journal of Statistical Software 76, 1–32 (2017). 64. [64].Kong, X. et al. Association of medical male circumcision and antiretroviral therapy scale-up with community HIV incidence in Rakai, Uganda. JAMA 316, 182–190 (2016). 65. [65].Ssempijja, V. et al. Results of early virologic monitoring may facilitate differentiated care monitoring strategies for clients on ART, Rakai, Uganda. Open Forum Infectious Diseases 5, ofy212 (2018). 66. [66].Ssempijja, V. et al. Adaptive viral load monitoring frequency to facilitate differentiated care: A modeling study from Rakai, Uganda. Clinical Infectious Diseases 71, 1017–1021 (2019). 67. [67].Grabowski, M. K. et al. Prevalence and predictors of persistent human immunodeficiency virus viremia and viral rebound after universal test and treat: A population-based study. Journal of Infectious Diseases 223, 1150–1160 (2021). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/infdis/jiab021&link_type=DOI) 68. [68].Dan, S. et al. Estimating fine age structure and time trends in human contact patterns from coarse contact data: The bayesian rate consistency model. PLOS Computational Biology 19, e1011191 (2023). 69. [69].Lewontin, R. C. Sex, lies, and social science. New York Review of Books 42, 24–29 (1995). 70. [70].Weinhardt, L. S., Forsyth, A. D., Carey, M. P., Jaworski, B. C. & Durant, L. E. Reliability and validity of self-report measures of HIV-related sexual behavior: progress since 1990 and recommendations for research and practice. Archives of sexual behavior 27, 155–180 (1998). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1023/A:1018682530519&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=9562899&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000072993900003&link_type=ISI) 71. [71].Gregson, S., Zhuwau, T., Ndlovu, J. & Nyamukapa, C. A. Methods to reduce social desirability bias in sex surveys in low-development settings: experience in Zimbabwe. Sexually transmitted diseases 29, 568–575 (2002). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1097/00007435-200210000-00002&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=12370523&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000178492000002&link_type=ISI) 72. [72].Kelly, C. A. et al. Using biomarkers to assess the validity of sexual behavior reporting across interview modes among young women in Kampala, Uganda. Studies in family planning 45, 43–58 (2014). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1111/j.1728-4465.2014.00375.x&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=24615574&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 73. [73].Gabry, J. & Češnovar, R. cmdstanr: R interface to ‘CmdStan’ (2020). Available at [https://mc-stan.org/users/interfaces/cmdstan](https://mc-stan.org/users/interfaces/cmdstan). 74. [74].Hastie, T. Generalized additive models R package (2020). Available at [https://cran.r-project.org/web/packages/gam/gam.pdf](https://cran.r-project.org/web/packages/gam/gam.pdf). 75. [75].Gall, A. et al. Universal amplification, next-generation sequencing, and assembly of HIV-1 genomes. Journal of Clinical Microbiology 50, 3838–3844 (2012). [Abstract/FREE Full Text](http://medrxiv.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNtIjtzOjU6InJlc2lkIjtzOjEwOiI1MC8xMi8zODM4IjtzOjQ6ImF0b20iO3M6NTA6Ii9tZWRyeGl2L2Vhcmx5LzIwMjMvMTEvMTMvMjAyMy4wMy4xNi4yMzI4NzM1MS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 76. [76].Bonsall, D. et al. ve-SEQ: Robust, unbiased enrichment for streamlined detection and whole-genome sequencing of HCV and other highly diverse pathogens. F1000Research 4, 1062 (2015). 77. [77].Bonsall, D. et al. A comprehensive genomics solution for hiv surveillance and clinical monitoring in low-income settings. Journal of Clinical Microbiology 58, e00382–20 (2020). 78. [78].Jenkins, F. et al. Validation of a hiv whole genome sequencing method for hiv drug resistance testing in an australian clinical microbiology laboratory. medRxiv (2023). [https://arxiv.org/abs/](https://arxiv.org/abs/) [https://www.medrxiv.org/content/early/2023/07/06/2023.07.05.23292232.full.pdf](https://www.medrxiv.org/content/early/2023/07/06/2023.07.05.23292232.full.pdf). 79. [79].Wymant, C. et al. Easy and accurate reconstruction of whole HIV genomes from short-read sequence data with shiver. Virus Evolution 4, vey007 (2018). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/ve/vey007&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=29876136&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 80. [80].Ratmann, O. et al. HIV-1 full-genome phylogenetics of generalized epidemics in Sub-Saharan Africa: Impact of missing nucleotide characters in next-generation sequences. AIDS Research and Human Retroviruses 33, 1083–1098 (2017). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1089/AID.2017.0061&link_type=DOI) 81. [81].Zhang, Y. et al. Evaluation of phylogenetic methods for inferring the direction of human immunodeficiency virus (HIV) transmission: HIV prevention trials network (HPTN) 052. Clinical Infectious Diseases: An Official Publication of the Infectious Diseases Society of America 72, 30–37 (2020). 82. [82].Xi, X. Bayesian methods for source attribution using HIV deep sequence data. h.D. thesis, Imperial College London (2022). 83. [83].Katoh, K., Misawa, K., Kuma, K. & Miyata, T. MAFFT: a novel method for rapid multiple sequence alignment based on fast fourier transform. Nucleic acids research 30, 3059–3066 (2002). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/nar/gkf436&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=12136088&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000177154300016&link_type=ISI) 84. [84].Nguyen, L., Schmidt, H. A., von Haeseler, A. & Minh, B. Q. IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Molecular Biology and Evolution 32, 268–274 (2014). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1093/molbev/msu300.&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=25371430&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 85. [85].Poon, A. F. et al. Dates of HIV infection can be estimated for seroprevalent patients by coalescent analysis of serial next-generation sequencing data. AIDS 25, 2019–2026 (2011). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1097/QAD.0b013e32834b643c&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=21832936&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) 86. [86].Cohen, M. S., Gay, C. L., Busch, M. P. & Hecht, F. M. The detection of acute HIV infection. Journal of Infectious Diseases 202, S270–S277 (2010). [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1086/655651&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=20846033&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2023%2F11%2F13%2F2023.03.16.23287351.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000281911000002&link_type=ISI) 87. [87].Pantazis, N. et al. Discriminating between premigration and postmigration HIV acquisition using surveillance data. Journal of Acquired Immune Deficiency Syndromes 88, 117–124 (2021). 88. [88].Bellan, S. E., Dushoff, J., Galvani, A. P. & Meyers, L. A. Reassessment of HIV-1 acute phase infectivity: Accounting for heterogeneity and study design with simulated cohorts. PLOS Medicine 12, e1001801 (2015). 89. [89].Monod, M. et al. Regularised b-splines projected gaussian process priors to estimate time-trends in age-specific COVID-19 deaths. Bayesian Analysis 18 (2023). 90. [90].Jacob, P. E., Murray, L. M., Holmes, C. C. & Robert, C. P. Better together? statistical learning in models made of modules. arXiv (2017). Preprint at [https://arxiv.org/abs/1708.08719](https://arxiv.org/abs/1708.08719). 91. [91].Wutzler, T. lognorm: Functions for the lognormal distribution. R package version 0.1 6 (2019). 92. [92].Monod, M. et al. Phylogenetic and epidemiologic data relating to age-specific HIV incidence and transmission in Rakai, Uganda, 2003-2018. (2023). Data at doi:10.5281/zenodo.8412741. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.5281/zenodo.8412741&link_type=DOI) [1]: /embed/inline-graphic-1.gif [2]: /embed/graphic-6.gif [3]: /embed/graphic-7.gif [4]: /embed/inline-graphic-2.gif [5]: /embed/inline-graphic-3.gif [6]: /embed/graphic-8.gif [7]: /embed/inline-graphic-4.gif [8]: /embed/inline-graphic-5.gif [9]: /embed/graphic-9.gif [10]: /embed/graphic-10.gif [11]: /embed/inline-graphic-6.gif [12]: /embed/inline-graphic-7.gif [13]: /embed/inline-graphic-8.gif [14]: /embed/inline-graphic-9.gif [15]: /embed/graphic-11.gif [16]: /embed/graphic-12.gif [17]: /embed/graphic-13.gif [18]: /embed/inline-graphic-10.gif [19]: /embed/inline-graphic-11.gif [20]: /embed/inline-graphic-12.gif [21]: /embed/inline-graphic-13.gif [22]: /embed/inline-graphic-14.gif [23]: /embed/inline-graphic-15.gif [24]: /embed/inline-graphic-16.gif [25]: /embed/inline-graphic-17.gif [26]: /embed/inline-graphic-18.gif [27]: /embed/inline-graphic-19.gif [28]: /embed/inline-graphic-20.gif [29]: /embed/inline-graphic-21.gif [30]: /embed/inline-graphic-22.gif [31]: /embed/inline-graphic-23.gif [32]: /embed/inline-graphic-24.gif [33]: /embed/inline-graphic-25.gif [34]: /embed/inline-graphic-26.gif [35]: /embed/inline-graphic-27.gif [36]: /embed/graphic-14.gif [37]: /embed/graphic-15.gif [38]: /embed/graphic-16.gif [39]: /embed/inline-graphic-28.gif [40]: /embed/inline-graphic-29.gif [41]: /embed/inline-graphic-30.gif [42]: /embed/inline-graphic-31.gif [43]: /embed/inline-graphic-32.gif [44]: /embed/inline-graphic-33.gif [45]: /embed/inline-graphic-34.gif [46]: /embed/graphic-17.gif [47]: /embed/graphic-18.gif [48]: /embed/graphic-19.gif [49]: /embed/graphic-20.gif [50]: /embed/inline-graphic-35.gif [51]: /embed/inline-graphic-36.gif [52]: /embed/inline-graphic-37.gif [53]: /embed/inline-graphic-38.gif [54]: /embed/graphic-21.gif [55]: /embed/graphic-22.gif [56]: /embed/graphic-23.gif [57]: /embed/graphic-24.gif [58]: /embed/graphic-25.gif [59]: /embed/graphic-26.gif [60]: /embed/graphic-27.gif [61]: /embed/graphic-28.gif [62]: /embed/graphic-29.gif [63]: /embed/inline-graphic-39.gif [64]: /embed/inline-graphic-40.gif [65]: /embed/inline-graphic-41.gif [66]: /embed/inline-graphic-42.gif [67]: /embed/graphic-30.gif [68]: /embed/inline-graphic-43.gif [69]: /embed/graphic-31.gif [70]: /embed/inline-graphic-44.gif [71]: /embed/inline-graphic-45.gif [72]: /embed/inline-graphic-46.gif [73]: /embed/inline-graphic-47.gif [74]: /embed/graphic-32.gif [75]: /embed/inline-graphic-48.gif [76]: /embed/inline-graphic-49.gif [77]: /embed/inline-graphic-50.gif [78]: /embed/inline-graphic-51.gif [79]: /embed/inline-graphic-52.gif [80]: /embed/inline-graphic-53.gif [81]: /embed/inline-graphic-54.gif [82]: /embed/graphic-33.gif [83]: /embed/graphic-34.gif [84]: /embed/graphic-35.gif [85]: /embed/graphic-36.gif [86]: /embed/graphic-37.gif [87]: /embed/graphic-38.gif